review - florida state universityweb2.physics.fsu.edu/~zhou/reprints/cr128.pdf · 2016. 1. 5. ·...

23
Subscriber access provided by UNIV OF CALIFORNIA CDL ACQUISITIONS Chemical Reviews is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Review Fundamental Aspects of Protein#Protein Association Kinetics G. Schreiber, G. Haran, and H.-X. Zhou Chem. Rev., 2009, 109 (3), 839-860• DOI: 10.1021/cr800373w • Publication Date (Web): 05 February 2009 Downloaded from http://pubs.acs.org on March 11, 2009 More About This Article Additional resources and features associated with this article are available within the HTML version: Supporting Information Access to high resolution figures Links to articles and content related to this article Copyright permission to reproduce figures and/or text from this article

Upload: others

Post on 27-Jan-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • Subscriber access provided by UNIV OF CALIFORNIA CDL ACQUISITIONS

    Chemical Reviews is published by the American Chemical Society. 1155 SixteenthStreet N.W., Washington, DC 20036

    Review

    Fundamental Aspects of Protein#Protein Association KineticsG. Schreiber, G. Haran, and H.-X. Zhou

    Chem. Rev., 2009, 109 (3), 839-860• DOI: 10.1021/cr800373w • Publication Date (Web): 05 February 2009

    Downloaded from http://pubs.acs.org on March 11, 2009

    More About This Article

    Additional resources and features associated with this article are available within the HTML version:

    • Supporting Information• Access to high resolution figures• Links to articles and content related to this article• Copyright permission to reproduce figures and/or text from this article

    http://pubs.acs.org/doi/full/10.1021/cr800373w

  • Fundamental Aspects of Protein-Protein Association Kinetics

    G. Schreiber,*,† G. Haran,*,‡ and H.-X. Zhou*,§

    Departments of Biological Chemistry and Chemical Physics, Weizmann Institute of Science, Rehovot, 76100, Israel and Department of Physics andInstitute of Molecular Biophysics, Florida State University, Tallahassee, Florida 32306

    Received May 22, 2008

    Contents

    1. Introduction 8391.1. Overview of Protein Association Kinetics 8411.2. Analytical Models of Protein Association Rate

    Constants842

    2. Methods for Studying Protein-Protein AssociationKinetics

    843

    2.1. Experimental Methods 8432.2. Computer Simulations and Modeling 843

    3. Intermediates along the Association Pathway 8453.1. Transient Complex 8453.2. Other Intermediates 8473.3. Mapping Intermediates by Mutations 848

    3.3.1. Φ-Value Analysis 8483.3.2. Double-Mutant Cycle Analysis 849

    4. Probing Electrostatic Enhancement of AssociationRates

    850

    4.1. Ionic Strength Dependences of ka and kd 8504.2. Probing the Contributions of Individual Residues

    Toward ka by Mutation851

    4.2.1. Altering Association Kinetics by ProteinDesign

    851

    4.2.2. Proteomics View of ‘Hotspots’ forAssociation

    852

    4.3. Prediction of Electrostatic Rate Enhancement bythe Transient-Complex Theory

    853

    5. Interaction Dynamics of Membrane-AnchoredProteins

    854

    6. Roles of the Depletion Effect and Crowding inAssociation Kinetics

    855

    6.1. Translational and Rotational Diffusion ofProteins in Concentrated Solutions and Cells

    855

    6.2. Depletion Effect 8566.3. Crowding and Protein-Protein Association

    Kinetics857

    6.3.1. Case of TEM-BLIP 8576.3.2. Surprises in Concentrated Solutions 858

    7. Conclusions 8588. Acknowledgments 8599. References 859

    1. IntroductionThe structure of a protein complex together with informa-

    tion about its affinity and other thermodynamic characteristicsprovide a “frozen” view of the complex. This picture ignoresthe kinetic nature of protein-protein association and dis-sociation, which is of major biological and biophysicalinterest. This review focuses on recent advances in decipher-ing the kinetic pathway of protein complex formation, thenature of the precomplex formed through diffusion (whichwe termed the “transient complex”1), the transition state, andother intermediates (such as the so-called encounter complex)along the association pathway.

    Protein-protein association is at the center of diversebiological processes ranging from enzyme catalysis/inhibitionto regulation of immune response by cytokines. The associa-tion rates often play a critical role in such processes, as insituations where speed is of essence.2 For example, the purplecone snail and other venomous animals capture prey withremarkable efficiency and speed by releasing toxins thatrapidly bind to ion channels;3 the green mamba achieves asimilar feat by targeting acetylcholinesterase (AChE), anenzyme essential for the integrity of neural transmission.4

    Bacteria such as Escherichia coli and Bacillus amylolique-faciens excrete nucleases as weapons against competitors orpredators. Defense of the producing cells from damage totheir own DNA or RNA by such nucleases requires rapidassociation with cognate inhibitors.5,6 Indeed, in the lastexample rapid association is such a priority that the inhibitorbarstar has a cluster of acidic residues that facilitate associa-tion with the nuclease barnase, even though the clusteredcharges reduce folding stability.7 In the ruminant gut RNase

    * To whom correspondence should be addressed. G.S.: phone, +972-8-9343249; fax, +972-8-9346095; e-mail, [email protected].: phone, +972-8-9342625; fax, +972-8-9342749; e-mail, [email protected]. H.-X.Z.: phone, 850-6451336; fax, 850-6447244; e-mail,[email protected].† Department of Biological Chemistry, Weizmann Institute of Science.‡ Department of Chemical Physics, Weizmann Institute of Science.§ Florida State University.

    Volume 109, Number 3

    10.1021/cr800373w CCC: $71.50 2009 American Chemical SocietyPublished on Web 02/05/2009

  • A is required for degrading accumulated RNA; potentialtoxicity of leaked nuclease is prevented by rapid associationwith a ribonuclease inhibitor.8,9

    Reorganization of the Actin cytoskeleton provides yetanother illustration of the importance of rapid proteinassociation. Reorganization is attained through Actin po-lymerization, which is nucleated by the Arp2/3 complex. Thelatter is activated by the Wiskott-Aldrich Syndrome protein(WASp), which in turn is released from the autoinhibitedstate by the Rho GTPase Cdc42.10 As Actin polymerizationis initiated with a nucleation process, the speed of upstreamsignaling has a critical impact on the rate of polymerformation. It is thus not surprising that high association rateconstants have been observed between partners along thesignaling pathway.11,12 The high association rate constantbetween Cdc42 and WASp has been found to be essentialfor the latter to stimulate Actin polymerization as another

    Rho GTPase sharing 70% sequence identity, TC10, with anidentical dissociation rate constant but a 1000-fold lowerassociation rate constant failed to stimulate Actin polymer-ization.11 The failure to stimulate Actin polymerization inpatients carrying mutant WAS genes is the root cause of theWiskott-Aldrich Syndrome.

    Several other compelling arguments can be made for thebiological roles of rapid protein association.13 (a) Fastassociation may enhance binding affinity. High affinity canalso be achieved through slow dissociation; however, forproteins involved in signaling slow dissociation is not anoption since it implies a long-lasting bound state, whicheffectively corresponds to a permanent off or on switch. Agood example for this is the binding of Ras to its naturalaffector Raf. This protein dissociates within a fraction of asecond but maintains an affinity in the nanomolar rangethrough fast association. Moreover, the difference betweenthe natural effector, Raf, and the non-natural effector, Ral,lies in their rates of association with Ras.14 Therefore, evenif not for a direct reason (such as in stimulation of Actinpolymerization), the affinity requirement alone may call forfast association. (b) Enzyme-substrate binding is a deter-mining factor for the overall turnover rate and becomes therate-limiting step for catalytically “perfect” enzymes. Substrate-binding rate constants of such enzymes reach 108 M-1 s-1

    and beyond, as found for the ribotoxin restrictocin and RNaseA.15,16 (c) When several proteins compete for the samereceptor or when one protein is faced with alternativepathways, kinetic control, not thermodynamic control, domi-nates in many cases; this is especially true when dissociationis slow. For example, during protein synthesis cognate andnoncognate aminoacyl-tRNA synthetases can potentiallycompete for the same tRNA. As an additional example,consider newly synthesized proteins, which potentially faceaggregation if not isolated by a chaperone. From the pointof view of kinetic control, it is easy to see why rapid bindingof denatured proteins to the chaperonin GroEL has beenobserved.17 (d) Differences in binding rate between relatedproteins may serve as an additional mechanism for specific-ity, as can be suggested for Rho GTPases Cdc42 and TC10and for Ras effectors Raf and Ral.

    The examples and arguments presented above suggest thatrapid binding is as important as high affinity in the proper

    Gideon Schreiber received his Doctorate in Biochemistry at the HebrewUniversity in Jerusalem in 1992. After a postdoctoral period at the MRCin Cambridge, U.K., at the laboratory of Professor Alan Fersht, he joinedthe Weizmann Institute of Science as Senior Scientist. Presently, he isan associate professor at the Department of Biological Chemistry at thesame institute. His research interest is in the investigation of protein-proteininteractions from the basic understanding of the mechanism to proteindesign. His work spans from bioinformatics and algorithm development,biophysical bench work, protein design and engineering, to applied biology.In addition he is a director of the Israel Structural Proteomic Center, whoseaim is to provide structures of proteins and protein complexes to thebiological community.

    Gilad Haran is currently an associate professor in the Department ofChemical Physics of the Weizmann Institute of Science. He did his Ph.D.work at the Weizmann Institute with Professors Ephraim Katchalsky-Katzirand Elisha Haas. He was then a postdoctoral fellow with Professor RobinHochstrasser at the University of Pennsylvania. His laboratory is usingsingle-molecule spectroscopy to study molecular dynamics, from proteinfolding and association to charge-transfer dynamics on metal surfaces.

    Huan-Xiang Zhou received his Ph.D. degree from Drexel University in1988. He did postdoctoral work at the NIH with Attila Szabo. After facultyappointments at HKUST and Drexel he moved to Florida State Universityin 2002. His group does theoretical, computational, and experimentalresearch on protein-protein interactions, crowding and confinement effectsof cellular environments, and biological roles of protein dynamics. He iscurrently the Editor-in-Chief of PMC Biophysics.

    840 Chemical Reviews, 2009, Vol. 109, No. 3 Schreiber et al.

  • functioning of proteins. It is now increasingly recognizedthat proteins function in the context of multicomponentcomplexes. Manipulating association rate constants of variouscomponents presents unique opportunities for the control ofprotein functions. Many interactions between proteins arealso targeted for drug development; in designing such drugs,both high affinity and rapid binding should be taken intoconsideration.

    1.1. Overview of Protein Association KineticsThe observed rate constants of protein association span a

    wide range from 109 M-1 s-1 (Figure 1). Incomprehending these values a basic fact is that for twoproteins to recognize each other their interfaces have to beoriented with high specificity. A relative rotation of as littleas a few degrees or a relative translation by a few Angstromsis sufficient to break all specific interactions between the twoproteins.18 The rate of association of a protein complex islimited by diffusion and geometric constraints of the bindingsites and may be further reduced by subsequent chemicalprocesses.19

    To better understand the kinetics of association of twoproteins (A and B) it is useful to consider the process asgoing through an intermediate state (A*B) in which the twoproteins have near-native separations and orientations.1,20-23

    We refer to this intermediate state as the transient complex,1,20

    noting that it is sometimes also termed the encountercomplex.24 A more detailed discussion of terminology as wellas the specification of the ensemble of configurations makingup the transient complex is provided in section 3. From thisensemble, conformational rearrangement can lead to thenative complex (C). Accordingly we have the kinetic scheme

    A+B {\}kD

    k-DA*B98

    kcC (1)

    While the first step of this scheme depends on relativediffusion between the protein molecules, the second step isakin to an intramolecular chemical reaction and can thereforebe described by the classical transition-state theory25 (withthe transition state located at the top of the free-energy barrierseparating A*B from C26) or by Kramers’ theory.27 The lattertheory accounts for barrier recrossing and models motionalong the reaction coordinate as diffusive. The overall rateconstant of association is

    ka )kDkc

    k-D + kc(2)

    which is bounded by the diffusion-controlled rate constant,kD, for reaching the transient complex. This limit is reachedwhen conformational rearrangement is fast relative to thedissociation of the transient complex (i.e., kc . k-D), leadingto

    ka ≈ kD (3a)

    In the opposite limit kc , k-D, conformational rearrangementor reaction becomes rate limiting and

    ka ≈ kckD ⁄ k-D ≡ kR (3b)

    Note that kD/k-D is the equilibrium constant for forming thetransient complex.

    There is no simple test that would place the associationof a given protein complex into the diffusion-controlled,reaction-controlled, or mixed regime. However, two generalstatements can be made. First, a diffusion-controlled rateconstant falls on the high end of the spectrum of observedvalues and a reaction-controlled rate constant falls on thelow end (Figure 1). Second, diffusion-controlled associationtypically involves only local conformational changes betweenthe unbound proteins and the native complex whereasreaction-controlled association typically involves gross changessuch as loop reorganization or domain movement. Thesepoints will be further amplified below. For later referencewe also introduce here the overall dissociation rate constant,kd, and the overall association constant, Ka ) ka/kd.

    In the transient complex the two protein molecules mustsatisfy translational/rotational constraints, which severelyhinder the diffusion-controlled rate constant kD. In theabsence of any biasing force, theoretical estimates put theresulting “basal” value, kD0, in the range from 105 to 106

    M-1 s-1 (see below).28-31 Antibody-protein association rateconstants are typically observed in this narrow range.32-34

    The value 105 M-1 s-1 thus may mark the start of thediffusion-controlled regime.

    To go beyond the basal rate constant kD0 and reachvalues in the range from 108 to 109 M-1 s-1 as observedfor many protein complexes3-6,8,9,15-17,35,36 (Figure 1),intermolecular forces must be present. For a force to speedup a “diffusion-controlled association” it must be presentin the diffusion process that leads to the transient complex.Indeed, analytical results on model systems show thatwhen the range of the force is reduced the resulting rateenhancement decreases drastically.29,37,38 For protein-proteinassociation the dominant long-range force is provided byelectrostatic interactions.

    Rate constants higher than kD0 are indeed usually correlatedwith favorable electrostatic interactions, as manifested bycomplementary charge distributions on the binding partners,which are illustrated in Figure 1 for four protein pairs. RNAand DNA have negative charges from phosphate groups;proteins targeting them (like barnase, RNase A, and DNaseE9) generally have enriched distributions of basic residueson the nucleic acid binding sites.39,40 Electrostatically en-hanced protein-nucleic acid association rates can thus beanticipated. The basic residues on the nucleic acid bindingsites can also be easily exploited by inhibitors (like barstar,ribonuclease inhibitor, and Im9): fast inhibition can beachieved through a concentration of acidic residues on thelatter molecules. Toxins blocking the Shaker potassium

    Figure 1. Wide spectrum of association rate constants. The redvertical line marks the start of the diffusion-controlled regime. Theshaded range marks the absence of long-range forces. (Adaptedwith permission from ref 1. Copyright 2008 Wiley Interscience.)

    Protein-Protein Association Kinetics Chemical Reviews, 2009, Vol. 109, No. 3 841

  • channel follow a similar strategy.23 Apparently for facilitatingthe conduction of the positively charged potassium ion themouth of the channel pore is lined with two rings of acidicresidues.41 To complement the resulting negative electrostaticsurface channel toxins have excess basic residues. Thecharge-complementarity argument can also be made torationalize the excess basic residues on fasciculin 2 (fas),the snake toxin targeting AChE. AChE uses a negativeelectrostatic surface around the entrance to the active-sitegorge for the fast binding of its positively charged substrate.4,42

    In other cases, such as in complexes formed by WASp andCdc4243 and by interleukin-4 (IL4) and its receptor,44 thereasons for a particular subunit to take up either positive ornegative charges are not obvious. That they nonetheless showcharge complementarity implicates functional roles of theirfast association.

    1.2. Analytical Models of Protein AssociationRate Constants

    The first theoretical result for the association rate constantwas obtained by Smoluchowski,45 who found that thediffusion-controlled rate constant for two uniformly reactingspheres to form a complex is

    kD0 ) 4πDR (4)

    where D is the relative translational diffusion constant andR is the contact distance between the centers of the twospheres. Debye recognized that the association rate betweenoppositely charged molecules can be increased by electro-static interactions.46 For two uniformly reacting spheres witha centrosymmetric interaction potential U(r) he found thediffusion-controlled rate constant to be

    kD ) 4πD ⁄∫R∞ eU(r)⁄kBTr-2 dr (5)where r is the intersphere distance and kBT is thermal energy.In the following discussion we will use kD0 to denote thediffusion-controlled rate constant in the absence of aninteraction potential (also referred to as the basal rateconstant) and use kD to denote the counterpart in the presenceof an interaction potential. The enhancement over the basalrate constant by an attractive interaction potential in theDebye model is quite modest. For example, for a Coulombicinteraction potential U(r) ) -Q/r one finds, upon evaluatingthe integral in eq 5, kD/4πDR ) (Q/kBTR)/(1 - e-Q/kBTR).The rate enhancement, kD/4πDR, is 9-fold when the mag-nitude of the potential is 9kBT at contact. This modest rateenhancement will be contrasted below with a much greaterenhancement predicted on a more realistic model for proteinassociation.

    For medium-sized proteins, R and D are on the order of40 Å and 20 Å2 ns-1, respectively. Equation 4 would predicta basal rate constant of 6 × 109 M-1 s-1. However, it isimportant to recognize that the stereospecific association oftwo proteins involves significant orientational constraints;thus, the Smoluchowski formula is of little use. Solc andStockmayer tackled the problem of orientational constraintsthrough a quasi-chemical approximation.47 For two sphereseach with a reactive patch they found the diffusion-controlledrate constant to be

    kD0 )4πDRF1F2

    Λ1Λ2 + [(1-Λ1)-1(1-Λ2)

    -1+(1-Λ1)

    -1(Λ2 -F2)-1 + (1-Λ2)

    -1(Λ1 -F1)-1]-1

    (6)

    where F1 and F2 are the surface fractions covered by thereactive patches. Relative diffusion has to bring theirseparation vector to within both reactive patches; therein thenative complex can form instantaneously. Specifying thereactive patches thus amounts to specifying the transientcomplex. An approximate expression, obtained by Berg,48

    for Λi (i ) 1 or 2), in the case of a patch spanning polarangles between 0 and δi is given by

    Λi ⁄ Fi )�i + cot(δi ⁄ 2)

    �i + sin(δi ⁄ 2)cos(δi ⁄ 2)(7)

    where �i ) [(1 + DiR2/D)/2]1/2 and Di is the rotationaldiffusion constant. For small patches one finds

    kD0 ⁄ 4πDR)F1�2 tan(δ2 ⁄ 2)+F2�1 tan(δ1 ⁄ 2) (8)

    The basal rate constant obtained by Brownian dynamicssimulations (see section 2.2)28 with simple structural as-sumptions about the transient complex (which was modeledby 2-3 correctly formed bonds between the proteins) is105-106 M-1 s-1, which is 4 orders of magnitude lower thanthe unrealistic Smoluchowski result. Such a rate constant ispredicted by eq 8 for reactive patches at sizes of δ1 ≈ δ2 ≈5° (with R and D at values quoted earlier and Di at 0.02ns-1). Results similar to eq 8 for the basal rate constant havebeen obtained by a number of different methods.49-52

    Reactive patches at sizes of ∼5° cover only a fraction of10-3 of the surface of each sphere. With the interspherevector having to lie within both reactive patches before thenative complex can form, the association rate constant isnaı̈vely expected to be lower by a factor of (103)2 ) 106. Asjust stated, the reduction factor predicted by eq 8 is 104. Thebasal rate constant thus appears to be 100-fold higher thannaı̈ve expectation. Northrup and Erickson28 explained thisapparent paradox by noting, based on Brownian dynamicssimulations, that when two proteins collide there is a highprobability that they will recollide several times beforeseparating again. Further, due to the separation of time scalesbetween rotational and translational diffusion they may rotatesignificantly between those collisions.

    Within the model of two spheres with reactive patchesthe influence of an interaction potential on the associationrate constant has also been studied.29,49 The expression forkD in the presence of a centrosymmetric potential U(r) is

    kD )4πDRe-U(R)⁄kBT[c1(0)c2(0)]

    2

    ∑l,l1,l2)0

    ∞ [ Fl(r) ⁄ r-dFl(r) ⁄ dr]r)R[c1(l1)c2(l2)]2Cll1l2(9)

    where Fl(r) satisfies

    r-2e�U(r)ddr

    r2e-�U(r)dFl(r)

    dr-

    [ l1(l1 + 1)D1 + l2(l2 + 1)D2D + l(l+ 1)r2 ]Fl(r)) 0 (10a)with the boundary condition Fl(r) f 0 as r f ∞, ci(li) aregiven by Legendre polynomials

    842 Chemical Reviews, 2009, Vol. 109, No. 3 Schreiber et al.

  • ci(li))Pli-1(cos δi)-Pli+1(cos δi) (10b)

    and Cll1l2 are given by Wigner 3 - j symbols

    Cll1l2 )2l+ 1

    (2l1 + 1)(2l2 + 1)(l1 l2 l0 0 0 ) (10c)Analytical results such as eq 9 can provide valuable insighton the properties of the association rate constant. Inparticular, in contrast to the Debye model in which rateenhancement by an attractive interaction potential is quitemodest, the rate enhancement in the more realistic patchmodel can be substantial. For reactive patches with sizes ofδ1 ≈ δ2 ≈ 5°, the rate enhancement is ∼5000-fold for apotential with a magnitude of 9kBT at contact.29 Moreimportantly, it is found that when U(r) is long ranged andδ1 and δ2 f 0, the dependence of kD on the interactionpotentialdisappearsexcept in theBoltzmannfactore-U(R)/kBT.29,49

    Therefore, the effect of the interaction potential can be simplycaptured by the Boltzmann factor. In section 2.2 we willreturn to this important result.

    In the following sections we give a detailed account oftheoretical and experimental studies of protein-proteinassociation kinetics. In section 2 we outline the mainexperimental and theoretical methods for studying associationkinetics. Section 3 presents experimental and theoreticalanalyses of the transient complex as well as the transitionstate for the final step in the association pathway. Section 4continues with the important subject of electrostatic rateenhancement. In section 5 we discuss protein-proteinassociation in the membrane environment, a field withgrowing interest. The subject of section 6 is the effect ofcrowding on association. Some final conclusions are drawnin section 7.

    2. Methods for Studying Protein-ProteinAssociation Kinetics

    The association kinetics between proteins has been studiedvigorously using both experimental and theoretical methods.There are some basic differences in studying binding ofproteins and their folding. In folding the unfolded state isnot well defined while the unbound state of proteins has awell-defined structure. This is not true in all cases as manyprotein-interaction domains were found in recent years tobe unstructured, and thus, binding and folding in those casesare coupled.53-56 However, here we review mostly thebinding of structured proteins. A second main differencebetween binding and folding is that folding is a first-orderprocess, where overall translational and rotational diffusionplays only a limited role. Conversely, as discussed in section1, for binding overall translational and rotational diffusionplay a major role in dictating the rate constant of the reaction.The existence of well-defined unbound structures simplifiesthe theoretical study of binding. For heterocomplexes thisalso simplifies the experimental study as one has a clearstarting point for the reaction, the mixture of the separatecomponents of the complex, and the possibility to follow inreal time complex formation. In this review we focus on theexperimental and theoretical work done to learn more aboutassociation between single globular proteins to formcomplexes.

    2.1. Experimental MethodsMost of the experimental work reviewed here was done

    with the proteins being in free solution using methods suchas stopped-flow spectrometry and NMR spectroscopy. Thisexcludes the results of many measurements done usingsurface plasmon resonance (SPR) with one of the proteinsattached to a surface and the other one in solution.57 Directcomparisons of association rate constants for a number ofprotein-protein complexes showed that SPR data mayprovide different ka values from those obtained in solution.58,59

    These differences may be attributed to mass transport,protein-immobilization effects, surface charges, crowding,etc.60,61 As these factors introduce additional complexity tothe analysis of protein association kinetics that is not directlyrelated to the association reaction, we review here onlysolution data.

    2.2. Computer Simulations and ModelingOne can model protein-protein association by simulat-

    ing the translational and rotational Brownian motion ofthe subunits. In these Brownian dynamics simulationsconformational fluctuations within the subunits are ne-glected, and hence, time scales far beyond those accessibleto molecular dynamics simulations can be explored.Brownian dynamics simulations make it possible tocalculate the diffusion-controlled rate constant kD forprotein shapes and interaction potentials beyond the scopeof analytical theories. This approach has been used bymany groups.24,28,49,62-71 In such a calculation one mustspecify a precise set of conditions, which when satisfiedsignifies formation of the native complex. This set ofconditions, typically implemented as an absorbing bound-ary in Brownian dynamics simulations, amounts to defin-ing the transient complex. Rather than being guided byany theoretical considerations, the location of the absorb-ing boundary is usually proposed in an ad hoc way andoften adjusted for best agreement with experiment. Forexample, in their Brownian dynamics simulations ofbarnase-barstar association Gabdoulline and Wade24tested three different specifications of an absorbing boundaryagainst experimental data. Two of these were based on atomdistances, and the third was based on electrostatic interactionenergy; each specification had an adjustable parameter.

    We will give a more detailed account on the specificationof the transient complex in section 3.1. Here we address theeffect of interaction potential, particularly one arising fromelectrostatic interactions between two associating proteins.In Brownian dynamics simulations the effect of electrostaticinteractions can be modeled by accounting for their influenceon the translational and rotational Brownian motion of theproteins. In principle, the electrostatic force and torque ontheproteinscanbecalculatedbysolvingthePoisson-Boltzmannequation

    ∇ ε(r) ∇ Φ(r))-4π[F(r)+M(r) ∑i:ions ciqie-qiΦ(r)⁄kBT]

    (11)

    where ε is the dielectric constant, Φ is the electrostaticpotential, F is the charge density inside the proteins, M is afunction with a value of 1 in all regions where mobile ionsin the solvent are accessible and 0 elsewhere, and qi and ciare the charges and bulk concentrations of mobile ions,respectively. The solution of the Poisson-Boltzmann equa-

    Protein-Protein Association Kinetics Chemical Reviews, 2009, Vol. 109, No. 3 843

  • tion is simplified by linearization leading to

    ∇ ε(r) ∇ Φ(r))-4πF(r)+M(r)εsκ2Φ(r) (12)

    where εs is the dielectric constant of the solvent and κ )(8πe2I/εskBT)1/2 is the Debye-Hückel screening parameterwith I ) ∑i: ionsciqi2 /2e2, the ionic strength. However, evenafter linearization, solving the Poisson-Boltzmann equationon the fly during a Brownian dynamics simulation isprohibitively expensive. One thus has had to rely onapproximations such as treating one of the proteins as a setof test charges62 (which leads to significant errors fromneglecting the low-dielectric region of the protein interior49)or the more elaborate effective-charge model.72 Unfortu-nately, the approximations are worst when the proteins arein close proximity, precisely where electrostatic interactionsare expected to have the strongest influence on kD.

    An algorithm is needed to obtain kD from the simulations.A widely used algorithm was developed by McCammon andco-workers.73 According to this algorithm one protein istranslationally immobilized at the origin and trajectories ofthe other protein are initiated uniformly on a spherical surface(at a radial distance r ) b) (see Figure 2a). This “b” surfaceshould be chosen far away from the immobilized protein suchthat the electrostatic interaction potential is centrosymmetric.If rotational motion is allowed, then the starting orientationsof the two proteins are random. The diffusion-controlled rateconstant is given by

    kD ) 4πDbf∞ (13)

    where f∞ is the fraction of trajectories that, instead of escapingto infinity, reach the absorbing boundary specifying theconditions for formation of the native complex. To find f∞,an outer absorbing boundary at r ) q > b, called the q

    surface, is introduced. The trajectories are terminated wheneither the absorbing boundary for complex formation or theq surface is reached. From the fraction, f, of trajectoriesterminated on the inner absorbing boundary one finds

    f∞ )f

    1- (1- f)kD(b) ⁄ kD(q)(14)

    where kD(b) and kD(q) are given by eq 5 with R replaced byb and q, respectively. It has been shown that eq 14 is strictlyvalid only when the q surface is far away from theimmobilized protein such that the equilibrium distributionof the mobile protein becomes isotropic.74

    An alternative algorithm75,76 has also been developed inwhich the absorbing boundary is extended into a “reactionregion” with a finite volume VRR (Figure 2b).18,77 A proteinpair that has reached the reaction region has a finite rate, γ,to form the native complex. In this treatment of proteinassociation the reaction rate γ models the conformationalrearrangement that brings the protein pair from the transientcomplex (specified by the reaction region) into the nativecomplex, that is, kc ) γ. The equilibrium constant kD/k-D isgiven by VRRe-〈U〉*/kBT, where 〈U〉* is the average interactionenergy within the transient complex. By starting Browniantrajectories from within the reaction region and obtainingthe surviving fraction, S, of the trajectories, the associationrate constant is obtained as ka ) γSVRRe-〈U〉*/kBT. Substitutingthe various results into eq 2, one finds that the diffusion-controlled association rate constant is given by

    kD ) γVRRe-〈U〉*⁄kBTS ⁄ (1- S) (15)

    While the surviving fraction S depends on the rate γ and onhow much the absorbing boundary is extended to form thereaction region, the result for kD given by eq 15 is onlydetermined by the absorbing boundary and the interactionenergy. We also note that the reaction-controlled associationrate constant (eq 3b) is

    kR ) kckD ⁄ k-D ) γVRRe-〈U〉*⁄kBT (16)

    in the present treatment.In a Brownian dynamics study of protein-protein associa-

    tion under the influence of electrostatic interactions it wasdiscovered that the survival fraction S is insensitive to thepresence of the electrostatic interaction energy.49 Then onecan write eq 15 as

    kD ) kD0e-〈U〉*⁄kBT (17)

    with kD0 as the basal rate constant given by

    kD0 ) γVRRS0 ⁄ (1- S0) (18)

    where S0 is the survival fraction without any biasing force.Note that a dependence on interaction energy as simple asthat given by eq 17 is expected for the reaction-controlledrate constant (see eq 16), but totally unexpected for thediffusion-controlled rate constant. Later an analytical deriva-tion of eq 17 was presented,78 which confirmed the tworequirements for the validity of eq 17 suggested in theoriginal work.49 These requirements are that the associationis stereospecific and the interaction energy is long ranged.Fortuitously (and fortunately!) these conditions are fulfilledby protein-protein association under the influence ofelectrostatic interactions. The accuracy of eq 17 has beendemonstrated against results from Brownian dynamicssimulations of protein-ligand binding20,79,80 and against

    Figure 2. Two algorithms for calculating the diffusion-controlledassociation rate from Brownian dynamics simulations. (a) Algorithmof Northrup et al.73 (b) Algorithm of Zhou.75 In either algorithmBrownian trajectories can be terminated by two triggers. In a thetriggers are absorption by either the inner absorbing boundary orthe q surface; in b the triggers are either reaction in the reactionregion or when the lifetime of the trajectory exceeds a cutoff.

    844 Chemical Reviews, 2009, Vol. 109, No. 3 Schreiber et al.

  • analytical results given by eq 9 for model systems.29 Thisequation resolves one of the two main obstacles to reliableprediction of protein association rate constants by making itpossible to rigorously treat electrostatic interactions. Theeffect of electrostatic interactions is captured by the Boltz-mann factor e-〈U〉*/kBT, which can be obtained by averagingover a relatively small number of representative configura-tions in the transient complex. The basal rate constant kD0still needs to be obtained through force-free Browniandynamics simulations, but these simulations are inexpensive.

    At first glance eq 17 looks like the transition-state theoryof the Eyring type.25 There is no relation between the two.As already emphasized, eq 17 is for a diffusion-controlledrate constant, but Eyring’s theory is for energy barrier-crossing processes that are activation controlled. Kramers27

    has shown that Eyring’s theory is a very poor approximationwhen barrier crossing becomes diffusion controlled. How-ever, there is no link between eq 17 and Kramers’ theoryfor diffusion-controlled barrier crossing either. In particular,the average interaction energy of the transient complex, 〈U〉*,which superficially resembles the energy barrier in Eyring’sand Kramers’ theories, is typically negative (leading to rateenhancement) instead of positive. Unlike the processesstudied by Eyring and Kramers, where the energy barrier iswhat hinders the rate, the diffusion-controlled rate of proteinassociation is hindered by the translational/rotational con-straints of the transient complex. Moreover, eq 17 does notrequire a reaction coordinate like the theories of Eyring andKramers. In fact, as will be discussed next, the specificationof the transient complex involves at least six degrees offreedom: three for translation and three for rotation.

    3. Intermediates along the Association PathwayThroughout this review we use the word “intermediate”

    as a generic term for a set of configurations in which thereis a degree of correlation in translation and rotation betweenthe associating proteins; it does not necessarily correspondto a local energy minimum or a species trapped in kineticexperiments. We refer to the latter as a kinetic intermediate.As discussed in section 1.1, the association process startswith formation of the transient complex (sometimes termedthe encounter complex, see section 3.1 below). The transientcomplex is close to the native complex in relative separationand relative orientation between the subunits but still missesa majority of the short-range interactions characterizing thenative complex. Formation of these short-range interactionsrequires passage through another fleeting structure, thetransition state for conformational rearrangement.26,81 Howcan structural information be obtained on these intermediates?

    Structural studies are routinely done to elucidate thetransition state and kinetic intermediates of protein foldingor enzyme catalysis. A range of experimental tools has beendeveloped for this task. NMR is a powerful tool to pin downthe residual structures of the unfolded state as well as capturetransient folding intermediates.82 Φ-Value analysis revealswhether specific interactions are formed already during akinetic intermediate or the transition state of the reaction.56,83,84

    Time-resolved spectroscopy and single-molecule spectros-copy are powerful tools which are frequently applied toinvestigate kinetic intermediates and transition states infolding.85-88 While these experimental tools provide only apartial view, they are extremely valuable for molecular

    dynamics simulations and other computational studies as theyprovide experimental reference points to benchmark simula-tions.89-91

    In comparison to studies on protein folding, structuralstudies on intermediates along the pathway of protein-proteinassociation are much less common. This may be partlyattributed to the technical difficulties stemming from the lowpopulation of the binding intermediates and the ill-definednature of the transition state for binding. Still, the develop-ment of protein-engineering tools, NMR, and time-resolvedoptical spectroscopy resulted in a number of interestingexperimental studies shedding light on the mechanisms ofprotein-protein association. Valuable guidance has also beenprovided by theory. In the following we focus on studiescharacterizing the transient complex and the transition stateof association.

    3.1. Transient ComplexIn theoretical predictions of association rate constants one

    must specify the transient complex. In different contexts thetransient complex has appeared as reactive patch, absorbingboundary, reaction region, and encounter complex. The adhoc way by which the transient complex is specified inBrownian dynamics studies is pointed out in section 2.2. Thesituation was no better when eq 17 was first used to predictassociation rate constants. In one early such study onbarnase-barstar association77 the transient-complex ensemblewas specified by adjusting the ranges of relative translationand rotation between the two proteins to match the experi-mental data at high ionic strength. Miyashita et al.92 directlyused experimental data for the association of cytochromec2 with the bacterial reaction center to locate the transient-complex ensemble in the 6-dimensional translation androtation space. The correlation coefficient between calculatedchanges in the electrostatic interaction energy by mutationsand the experimental counterpart

    ∆∆G‡-U ≡-kBT ln(kamut ⁄ ka

    wt) (19)

    was used as an effective energy in a Monte Carlo simulation.Schreiber and co-workers93,94 further simplified eq 17 bycalculating 〈U〉* from an empirical function simply on thenative complex, thus avoiding the specification of thetransient complex; the empirical function was parametrizedon experimental data. Applications of that approach arepresented in section 4.2.

    In order to predict association rate constants from theoryalone, the transient-complex ensemble has to be specifiedwithout reference to experiment. A solution to this chal-lenging problem was proposed in a recent paper18 based onanalyzing the interaction energy landscape of associatingproteins.

    In a complete theory the overall association rate constantka should not be sensitive to where the transient complex isplaced. If it is placed far away from the native complex,then kD will be large but kc will be small. Conversely, if itis placed very close to the native complex, then kD will bereduced but kc will become very large. Either way, eq 2 isexpected to give nearly the same result for ka (as explicitlyshown for a model system95). However, given the consider-able difficulty and uncertainty in the calculation of kc it ishighly desirable to use kD as a close approximation for ka.Then there is an optimal location for placing the transientcomplex.22 If it is placed too far from the native complex,

    Protein-Protein Association Kinetics Chemical Reviews, 2009, Vol. 109, No. 3 845

  • then the resulting kD would not be a useful approximationfor ka. On the other hand, placing the transient complex tooclose to the native complex would mean that short-rangeinteractions and conformational rearrangement have to bedealt with in calculating kD. The native complex sits in adeep well in the interaction energy landscape.18 The optimalplacement for the transient-complex ensemble is at the outerboundary of the native-complex energy well (Figure 3a).18,22

    The algorithm for identifying the transient complex wasbased on the following observation: inside the native-complex energy well translation and rotation are restricted,but once outside, the proteins gain significant translational/rotational freedom.18 Thus, the outer boundary of the native-complex energy well coincides with the onset of translational/rotational freedom. To simplify the calculations required fordetermining the transient complex, the short-range interactionenergy stabilizing the native complex was modeled by thenumber of contacts, Nc, formed between the protein partners.Translational/rotational freedom was measured by σ�(Nc), thestandard deviation of the rotation angle � in configurationswith a given contact level Nc. A sharp increase in σ� withdecreasing Nc marks the onset of translational and rotational

    freedom and hence the location of the transition complex(Figure 3b). As illustrated in Figure 3c for the barnase-barstarpair, the resulting model for the transient complex is acompact ensemble of configurations which marks the onsetof a sharp increase in translational and rotational freedomas the protein partners move out of the native-complex energywell. In the ensemble of transient-complex configurationsof the barnase-barstar pair the displacement vector r ismostly confined in a 4 × 4 × 1 Å3 volume, correspondingto distances of 4.5 ( 0.5 Å between the protein surfaces(allowing for at least one layer of solvent). Within thistranslational volume the relative orientations of the twoproteins are also severely restricted with the rotation angle� mostly confined in the 0°-10° range.

    Since kD is used as the prediction for ka, eq 17 can berewritten as an equation for ka

    ka ) ka0e-〈Uel〉*⁄kBT (20)

    where ka0 is the rate constant for reaching the transientcomplex in the absence of any biasing force and can beobtained by Brownian dynamics simulations, and the interac-tion energy is now denoted with a subscript “el” to signify

    Figure 3. Location of the transient complex within the interaction energy landscape. (a) The energy landscape of protein-protein association.Translation (r) and rotation (e, �) coordinates lie in the horizontal plane. The transient-complex ensemble is indicated by a green ring. Thesmooth energy landscape results from keeping the two proteins in their native internal conformations. (b) Identification of the transientcomplex, consisting of configurations with the contact level Nc*, where σ� is poised for a sharp increase with decreasing Nc. (c) Translationalvolume of the transient complex, which is the projection of the hypersurface Nc ) Nc* into the 3 dimensions of translation, for thebarnase-barstar pair. A cluster of red dots represent the locations of the displacement vector r (yellow arrow). (Parts a and b were adaptedwith permission from ref 1. Copyright 2008 Wiley Interscience.)

    846 Chemical Reviews, 2009, Vol. 109, No. 3 Schreiber et al.

  • that only electrostatic interactions are to be included. Theneglect of short-ranged nonelectrostatic effects from theBoltzmann factor is based on two considerations. First, asFigure 3c shows, the transient-complex configurations areseparated by at least one layer of solvent; therefore, short-ranged forces such as hydrophobic and van der Waalsinteractions are relatively weak in the diffusion process,leading to the transient complex. Second, short-range interac-tions, even when present within the transient complex,contribute much less to rate enhancement (i.e., ka/ka0)compared to long-range interactions. However, short-rangeinteractions are essential for determining the location andsize of the transient-complex ensemble in configurationalspace, which in turn affect the magnitude of ka0. A transient-complex ensemble that is less restricted in translation androtation will lead to a higher ka0. Variation of the restrictionin translation and rotation within the transient complex withsolvent conditions or among different protein complexes canbe viewed as a configurational entropy effect. The basal rateconstant ka0 captures all the contributions of short-rangeinteractions and configurational entropy.

    The structural model for the transient-complex ensemblepresented above along with eq 20 constitutes the transient-complex theory for predicting protein-protein associationrate constants. With this theory the two main obstacles facedby the traditional approach of Brownian dynamics simula-tions, the necessity for approximate treatment of electrostaticinteractions and the ad hoc specification of the transientcomplex, are both resolved. Applications of this theory arepresented in section 4.3.

    3.2. Other IntermediatesThe term “encounter complex” has frequently been used

    in describing the pathway of protein-protein association. Itappears that this term is assigned different meanings indifferent contexts. For example, in calculating associationrate constants by Brownian dynamics simulations, Gab-doulline and Wade used this term to refer to “the end-pointof diffusional association”,24,96 which would be similar towhat we have defined as the transient complex. However,these same authors later also used encounter complex to referto low free-energy regions in configurational space.70 In theparamagnetic NMR experiments described next, encountercomplex refers to a minor, dynamic state that is in equilib-rium with a dominant, stereospecific complex.

    Using the paramagnetic relaxation enhancement (PRE)NMR technique on the electron-transfer pair of yeastcytochrome c peroxidase (CcP) and iso-1-cytochrome c (Cc)Volkov et al.97 determined the structure of the dominant,well-defined complex and delineated the configurationalspace of the minor, dynamic state. The latter, as just noted,was referred to as encounter complex. The dominant complexis very similar to the X-ray structure of the pair and occupiedfor >70% of the time. In the encounter complex, Cc occupiesa region around the position in the dominant complex; thenegative charges of CcP bordering this region suggest “thatelectrostatic attraction plays a dominant role in determiningthe nature of the encounter complex”. However, the data alsoseem to indicate that the encounter complex containsconfigurations in which Cc has its “back” side facing CcP.

    Tang et al. carried out a similar study for the associationbetween the phosphocarrier protein, Hpr, and three proteinsin the bacterial phosphotransferase system.98 They found arather diffuse encounter complex that qualitatively correlates

    with the electrostatic surface potentials on the interactingproteins. These experiments were performed in the absenceof salt, in which case nonspecific electrostatic attraction ofthese highly charged proteins is strong over a long range.6

    In a follow-up paper99 Tang et al. showed that the nonspecificpart of the encounter complex is reduced to a large extentby adding salt, while more specific configurations (locatednear the region of the native complex) were less affected.This study seems to suggest that not all regions of theencounter complex lead to productive association; it is theregion which does lead to productive association that wespecifically refer to as the transient complex.

    Spaar et al.70 developed a method for mapping theencounter complex along the association path from BDsimulations. Applying this method to the barnase-barstarinteraction they suggest two minima along the pathway, oneleading (according to an analysis of successful trajectories)to association, while the other does not. Still, they suggestthat this second region may be helpful to steer barstar intothe region of the transient complex. In a yet unpublishedstudy combining simulation and experiment Spaar et al.aimed to directly evaluate the importance of nonspecificencounter regions to association. Using the method developedby Spaar et al.70 on TEM1-BLIP association, the BDsimulations identified two main regions with low interactionenergies for the wild-type complex; however, most trajec-tories started at these regions did not lead to productivebinding. It was found that only mutations designed toincrease the size and energy of the encounter region nearthe binding interface had a major effect on the measuredassociation rate constant. Conversely, mutations increasingor decreasing the occupancy and energy of other intermedi-ates had no effect on the association rate constant. Theseresults indicate that not only the occupancy or energy of theencounter regions but also their location are important indetermining their effect on the association rate constant.Thus, some encounter regions are fruitful while others arefutile. This conclusion was in fact anticipated when theencounter region that does lead to productive association wassingled out to define the transient complex (see abovediscussion illustrated by Figure 3a).1

    In some cases, kinetic intermediates in binding have beenobserved. One such case is the interaction between colicinendonucleases and their cognate and noncognate immuno-proteins.100 Stopped-flow spectrofluorimetry shows that bind-ing of both Im9 and Im2 to E9 result in a rapid fluorescencequenching step (pre-equilibrium) with a concentration-dependent rate of formation followed by a second, concen-tration-independent slow rate (on a time scale of seconds).The data are best explained by a mechanism where cognateand noncognate complexes alike form a “dynamic encountercomplex”. Thereafter the cognate pair transitions to a high-affinity, stereospecific complex, whereas the noncognate pairremains dynamic without achieving high affinity.

    A somewhat different mechanism involving a pre-equi-librium complex followed by a reorganization step wassuggested for cystatin A binding papain.101,102 The biphasicbehavior of cystatin A/papain interaction was observed byfollowing independently the inhibition of the catalytic activityof papain (where a linear dependence of kobs on proteinconcentration was observed), while the spectroscopic probe(which monitors the accumulation of the native complex)showed a hyperbolic dependence of kobs on protein concen-

    Protein-Protein Association Kinetics Chemical Reviews, 2009, Vol. 109, No. 3 847

  • tration. This example demonstrates that characterization ofthe association behavior may depend on the probe used tomonitor it.

    3.3. Mapping Intermediates by Mutations3.3.1. Φ-Value Analysis

    Φ-Value analysis was successfully applied to map thetransition state for protein folding and became the goldenstandard for many simulations.103 It has been demonstratedthat this analysis can be used also for studying the rate-determining state (denoted by ‡) for protein-proteinassociation.14,56,104-106 The basic formulation for such analysisfor binding is given by

    Φa )∆∆G‡-U ⁄ ∆∆GC-U (21)

    where ∆∆G‡-U is defined by eq 19 and ∆∆GC-U isanalogously defined but with the association rate constantka replaced by the association constant Ka. Mutations thatinduce a similar effect on the rate-determining state (‡) andthe free energy of binding will have a Φ value of 1, whilemutations that have no effect on the rate-determining state(‡) but change the association constant will have a Φ valueof zero. The interpretation of this form of Φ-value analysisis particularly transparent when conformational rearrange-ment is rate limiting (cf. eq 3b). In this case ∆∆G‡-Urepresents the mutational effect on the free energy forforming the transition state between the transient complexand the native complex, and the Φ-value analysis thus reportson the transition state. If, on the other hand, conformationalrearrangement is fast (i.e., kc . k-D) then the analysis reportson formation of the transient complex.

    Figure 4 shows a Φ-value analysis for a large number of

    ‘hotspot’ mutations (those with ∆∆GC-U > 2 kcal/mol inmagnitude) collected from TEM1-BLIP, barnase-barstar,and Ras-Ral binding.6,14,93,107 The mutations were dividedinto three groups: one consisting of noncharged residues, thesecond consisting of charged residues located within thebinding interface, and the third consisting of charged residueslocated outside the binding interface. The reason we showonly ‘hotspot’ residues is to avoid erroneous Φ values asthe experimental error for ∆∆G measurements is on the orderof 0.3 kcal/mol. As only very few residues located outsidethe binding interface pass this criteria, we included in thisgroup also multiple mutations that were designed to specif-ically increase the association constant.14,93 The data clearlydemonstrate that noncharged mutations always have Φ valuesclose to zero, while charged residues located outside thephysical binding site have Φ values close to one. Chargedresidues located within the binding site have mixed values.A simple explanation of these Φ values is that they reporton the transient complexes of the three protein pairs. As notedin the discussion below eq 20, the transient complex ispredominantly stabilized by long-range electrostatic interac-tions with nonpolar interactions playing a minor role. Astriking result of the Φ-value analysis is that hydrophobicresidues appear to play only a minor role in formation ofthe transient complex. As most mutations were made to Ala,which retains a hydrophobic surface, one cannot rule outthe possibility that nonspecific hydrophobic interactions areformed at the transient complex also for the mutant, reducingthe Φ value. However, this contribution should be verylimited as mutations to Ala create cavities on the surfacethat are not fully filled up by the incoming protein.

    A hallmark of a partially reaction-limited reaction (i.e.,not purely diffusion controlled) is a nonlinear relationbetween the rate of association and the concentrations ofreactants. This has been observed for enzyme-substrateinteractions but rarely for protein-protein association.However, in the case of the association between RalGDS-RBD and Ras this nonlinearity could be experimentallydemonstrated, presenting direct evidence for the transientcomplex as a kinetic intermediate for association.14 Thenonlinearity was observed despite a high ka value (107 M-1

    s-1) and a putative basal rate of 6 × 105 M-1 s-1, whichshould in principle place this association in the diffusion-controlled regime (see section 1.1). Using the program PARE(see section 4.2.1),93 mutations of charged residues locatedat the periphery of the binding interface were designed, whichspecifically increased ka and did not affect kd. Stopped-flowmeasurements showed that the increase in ka of a +7 mutantwas a result of an increased rate of formation of the transientcomplex, while the rate of conversion to the native complexwas unchanged at a value of ∼400 s-1.14 From these data Φvalues of 1.0 and 0.9 for the transient complex and transitionstate, respectively, were calculated. This demonstrates thatfavorable electrostatic interactions introduced by mutationstabilize the transient complex and transition state to the sameextent as they stabilize the native complex. This study alsopoints to the difficulty in assigning a particular associationprocess as diffusion limited or reaction limited. At lowprotein concentrations the overall association rate may appearto be diffusion limited, but at high protein concentrations itmay appear to be reaction limited.

    The +7 RalGDS-RBD variant (obtained by electrostaticdesign) binds Ras with an association constant similar to thatmeasured for the native effector, Raf. Interestingly, the

    Figure 4. Φ-Value analysis of the transient complex/transition statefor association determined for ‘hotspot’ mutations (affecting thebinding affinity by >2 kcal/mol) collected from TEM1-BLIP,barnase-barstar, and Ras-Ral association.6,14,93 Φ values close to1 suggest that the involved residues have similar interactions inthe transient complex/transition state and in the native complex,while values close to 0 indicate the residues do not form anyinterprotein contacts in the transient complex/transition state.

    848 Chemical Reviews, 2009, Vol. 109, No. 3 Schreiber et al.

  • electrostatic potential map of the RalGDS-RBD variant isalso similar to that of Raf (Figure 5a) despite the verydifferent sequences of the two effectors (

  • for barnase-barstar and the complex between P. laminosumCyt fandplastocyaninsomebutnotallpairwisecharge-chargeinteractions were maintained, suggesting that the structuralspecificity of the transient complex/transition state is pre-served even at high salt but its size may shrink.119,120

    The experimental mutant and double-mutant cycle datameasured on the association rate constants were further usedto model the structures of the transient complex/transitionstate. In the study of Harel et al. the transient complex/transition state was modeled from the experimental ∆∆Gintq-Uvalues by searching for those interprotein orientations thatbest account for the experimental ∆∆G‡int values (Figure 6).Similarly, Miyashita92 related the experimental ka values ofmutant proteins to differences in the calculated electrostaticenergies for a wide range of cytochrome C2-reaction center(Cyt-RC) configurations. Both studies gave a very similardescription of the transient complex/transition state. In bothcases, the transient complex/transition state was stabilizedby electrostatic interactions with the ensemble of configura-tions spread out around the native complex, but in neithercase were short-range interactions formed, suggesting asolvated transient complex/transition state. The averagestructure of the transient complex/transition state was notnecessarily located exactly at the center of the bindinginterface but could be shifted toward one side of the interface.This was observed for both the Cyt-RC and the electrostati-callyoptimizedTEM1-BLIPpairs,whileforthebarnase-barsterpair the average structure of the transient complex/transitionstate overlapped the native complex.92,94,118 These charac-teristics bear striking resemblance to the compact structuralmodel of the transient complex derived from theory (section3.1). They suggest a certain pathway for association (downan energy funnel), which would help in speeding up theassociation.18 Contrary to these results, no indication for atransient complex with specific structures was found for theassociation between wild-type TEM1 and BLIP or betweenIFNR2 and IFNAR2 (Figure 6). Therefore, a diffusivetransient complex/transition state was suggested for theseprotein pairs. What distinguishes a diffusive transientcomplex from a compact one is the absence of definedinterprotein orientations. As was clearly shown for theTEM1-BLIP complex, mutations introduced through ratio-

    nal design can change the transient complex/transition statefrom diffusive to compact and vice versa.118

    All the evidence described above suggest the existence ofa transient complex/transition state along the associationpathway. The main points presented are (1) the binding ofRas to either Ral or Raf is biphasic with the first phase beingdiffusion limited and the second phase reaction limited, (2)increasing the electrostatic complementarity between Ras andRal specifically affects the rate of the first phase and doesnot affect the second phase of association, (3) increasingelectrostatic complementarity affects only the association butnot dissociation rate constant, (4) double-mutant cycleexperiments as well as Φ-value analyses clearly show thatonly charged residue pairs form interactions in the transientcomplex, while short-range interactions are largely unformedin the transient complex. Further formation of such short-range interactions, which are important for stabilizing thenative complex, may encounter a transition state. In the caseof a compact transient complex the transient complex alreadyhas interprotein orientations similar to those in the nativecomplex. The transition state will at least inherit such anorientational similarity. Therefore, in terms of gross structure,the transient complex and transition state seem to be close.

    4. Probing Electrostatic Enhancement ofAssociation Rates

    As discussed in previous sections, electrostatic interactionshave long been recognized as the dominant factor used byNature in order to enhance protein association beyond thebasal rates dictated by diffusion.13,49,121 In this section wediscuss experimental and theoretical methods to probe andanalyze the effect of electrostatic attraction on associationkinetics as well as protein engineering studies which enableus to modify and control this effect. The preceding sectionshave already presented a glimpse into some of these issues.

    4.1. Ionic Strength Dependences of ka and kdThe magnitude of the electrostatic attraction between two

    proteins can be modulated most simply by changing the ionicstrength of the solution. It has been recognized that whenthe association between two proteins is diffusion controlledmodulation of the electrostatic attraction leads to a universalphenomenon related to the association and dissociation rateconstants ka and kd, which show very disparate dependenceson ionic strength.22,122 The association rate constant decreasessignificantly with increasing ionic strength, whereas thedissociation rate constant is only modestly affected by ionicstrength. The disparate ionic strength effects have beenobserved on a number of the protein-protein complexesdiscussed above.4,5,11,35,109 Many other proteins conform tothe same behavior.123-133

    The compact structural model of the transient complexderived from theory, presented in section 3.1, provides a niceexplanation for the disparate effects of ionic strength onassociation and dissociation rate constants when associationis diffusion controlled.22,122 As the transient complex lies atthe outer boundary of the interaction energy well and henceis close to the native complex, ionic strength is expected toscreen electrostatic interactions in the two types of complexesto nearly the same extent. Hence, the association constantand association rate constant are expected to have nearly thesame dependence on ionic strength and the dissociation ratewould be little affected by ionic strength.

    Figure 6. Mapping the transient complex/transition state forprotein-protein association for the TEM1-BLIP and barnase-barstarcomplexes using double-mutant cycle data as constraints. Each pointrepresents the center of mass of one of the 2220 configurations ofTEM1 or barnase perturbed from the native complex (the mobileproteins in the simulations). BLIP and barstar are the fixed proteinsand are represented as ribbon. The point in the middle of each caprepresents the X-ray structure of the native complex. The differentcolors represent configurations selected by different filtering cutoffs;colder colors designate a configuration that passes a more stringentcutoff (thus, has a higher probability of occupancy in the transientcomplex/transition state). The TEM1-BLIP complex was electro-statically optimized using the program PARE by introducingmutations located outside the binding interface.

    850 Chemical Reviews, 2009, Vol. 109, No. 3 Schreiber et al.

  • The relation between ionic strength and ka can bequantitatively predicted from the solution of the Poisson-Boltzmann equation (eq 12) as will be discussed in section4.3. Empirically, it has been shown that this relation can bedescribed quite well by a Debye-Hückel-like approxi-mation77,93

    ln ka ) ln ka0 -U0 ⁄ kBT

    1+ κa(23)

    where U0 and a are fitting parameters. Comparison of eq 23with eq 20 shows that U corresponds to the electrostaticinteraction energy of the transient complex at κ ) 0 (i.e., inthe absence of salt). According to eq 23, a plot of ln ka versus(1 + κa)-1 is linear. The intercept of the line at (1 + κa)-1) 0 is ln ka0, from which the basal rate constant ka0 can bedetermined (Figure 7). The y value at (1 + κa)-1 ) 1 is lnka in the absence of salt, wherein the electrostatic attractionis maximized.134 This linear relation was shown to hold forthe association of the TEM1-BLIP, interferon-receptor,hirudin-thrombin, and barnase-barstar pairs and a het-erodimericleucinezipperforallsaltconcentrationstested.93,129,135

    Figure 7 clearly shows that electrostatic attraction maybe very strong at low salt. This phenomenon is well knownfor proteins and the basis of ion-exchange purification. Toavoid strong, nonspecific electrostatic attraction the ionicstrength in the cell is ∼150 mM. At this ionic strengthcharge-charge interactions are partially shielded, reducingthe negative effect of nonspecific interactions by allowingfor specific rate enhancement. A good example for this wasreported for the complex of barnase-barstar in the presenceof the poly ion hirudin.136 The association rate at low saltwas actually lower than in higher salt due to nonspecificinteractions of barnase with hirudin, which effectivelyremoved free barnase from the system. The rate peaked at150-200 mM salt and slowed down at higher salt (this timedue to masking of the charges). Thus, the physiological ionicstrength is optimal to obtain fast specific binding yet reducenonspecific binding.

    4.2. Probing the Contributions of IndividualResidues Toward ka by Mutation

    Measuring the effects of mutations on the rate constantsof association is a powerful tool to decipher the mechanismof association. As already discussed in section 3.2, extensivesite-directedmutagenesisofsurfaceresiduesoftheTEM1-BLIP,barnase-barstar, interferon-receptor, growth-hormone-receptor, IL4-receptor, and other protein pairs hasdemonstrated that mutations involving charged residues havethe largest effect on ka, while mutations of uncharged residesare much less important.6,93,107,108,135,137-141 This point isfurther illustrated in Figure 8, which presents the analysison the effects on ka and kd by a set of 55 single mutationslocated in the binding interfaces of the TEM1-BLIP,barnase-barstar, and interferon-receptor pairs. Except forthe A19W mutation in interferon, none of the neutralmutations had a large effect on ka, while many of them hadmajor effects on kd. This points toward a clear distinctionbetween the process of association and dissociation. Hotspotresidues affecting dissociation rates are both of charged anduncharged nature, while hotspots for association are almostalways charged residues. Moreover, the energy gained orlost though a mutation is of a different scale, while for ka itis rare to find a residue that alters the rate by over 10-fold;for kd it is common that >5 such residues are found withinan interface. While charge is important for association, themagnitude of perturbation of ka is not simply a measure ofthe change in charge but rather relates to the specific locationof the mutation and its contribution to the electrostatic energyof interaction between the two proteins in the transientcomplex (eq 20).1,13,77,93,94 For example, charged mutationslocated far away from the binding interface will not affectassociation even if these mutations are drastic.6

    4.2.1. Altering Association Kinetics by Protein Design

    As will be discussed in section 4.3, eq 20 allows for arigorous prediction of the association rate constant. Theprediction requires generation of the transient complex. For

    Figure 7. Association rate constants of wild-type and mutantTEM1-BLIP complexes determined at different salt concentrationswith ln ka plotted against (1 + κa)-1. The data can be fitted to aline according to eq 23. Figure 8. Changes in the rate constants of (a) association and (b)

    dissociation plotted against the change in charge resulting from amutation. The rate constants were measured for 55 mutations inthe barnase-barster, TEM1-BLIP, and IFNR2-IFNAR2 pairs.

    Protein-Protein Association Kinetics Chemical Reviews, 2009, Vol. 109, No. 3 851

  • design purposes a simplified approach is called for. In thesimplified approach taken by Schreiber and co-workers theelectrostatic interaction energy 〈Uel〉* was calculated on thenative complex itself93,94

    〈Uel 〉 * ≈ UC -UA -UB (24)Here UX, X ) A, B, or C, refers to the electrostatic energyof two proteins or their native complex given by

    UX )12∑i,j

    qiqjεsrij

    e-κ(rij-a)

    1+ κa(25)

    where qi are the charges of the atoms in a protein or theprotein complex and rij are the distances between the charges.The parameter a is the same as the one appearing in eq 23,and a value of 5.6 Å was found experimentally to give thebest results in the calculations.93,94 Despite their simplicityand the use of the native complex instead of the transientcomplex, eqs 24 and 25 were shown to give good estimatesof electrostatic contributions to the rate constants of associa-tion and of the way mutations will affect the rate constantsfor diverse systems such as RNaseA-RI, AChE-fas,TEM1-BLIP, hiruding-thrombin, CheY-CheA, Ras-Raf,and Ras-Ral (Figure 5b).9,14,93,94,132,142 This was found tobe true whether mutations were placed within or outside thebinding interfaces. Most informative were the cases wherecharged mutations were engineered outside the bindinginterfaces of protein-protein complexes, such as for theTEM1-BLIP and Ras-Ral pairs.14,93 In these cases, strongincreases in the rate constants of association were achieved(250- and 17-fold, respectively), in agreement with thecalculated values. These increased ka values were howevernot accompanied by a change in kd, leading to increasedassociation constants of the magnitudes described. Thisobservation has far-reaching implications for our understand-ing of the transient complex and transition state for associa-tion, as will be described below. It is important to note thateqs 24 and 25 successfully predict the changes in associationrate constant also for mutations located within the bindinginterfaces of protein-protein complexes, as shown for theassociation of the barnase-barstar, TEM1-BLIP, Ras-Ral,AChE-fas, hirudin-thrombin, CheY-CheA, and otherprotein pairs.6,132

    On the basis of eqs 24 and 25, two web servers forcalculating the changes in ka upon mutations were set up:PARE (http://www.weizmann.ac.il/home/bcges/PARE.html)and HyPare (http://bip.weizmann.ac.il/HyPare). The firstprovides a more exact calculation for single mutations, whilethe second provides a full analysis of the entire proteinsurface, including the automated identification of “hotspots”for association.94,143

    According to eq 20, basal rate constants are the rateconstants of association in the absence of electrostaticattraction. These can be calculated either by extrapolatingthe values of ka to the limit of infinite salt concentration orby introducing mutations that reduce the electrostatic interac-tion energy to zero. From such calculations the basal rateconstants were found to be 4.4 × 104 or 4 × 104 M-1 s-1for thrombin-hirudin association, 4.2 × 104 or 2.3 × 104M-1 s-1 for TEM1-BLIP association, and 1.4 × 107 or 1.5× 105 M-1 s-1 for barnase-barstar association (the twovalues are extrapolated from mutant or salt data, respectively,see Figure 7 and eq 23). Similarly, the values of ka0 were6.6 × 105 M-1 s-1 for Ras-Raf association and 2 × 105M-1 s-1 for AChE-fas association using mutation data.

    Except for barnase-barstar association, the basal rateconstants were the same whether calculated from salt ormutant data. The reason for the large difference between thetwo methods in predicting the basal rate constant forbarnase-barstar association is currently not clear;143 possiblythis is an indication of the limitation of the simplification ineqs 24 and 25 (see section 4.3 for further discussion).

    It is important to note that eq 25 does not treat explicitlythe contributions of noncharged residues to ka. While indeedthe contributions are small, they were found to be significantin a number of cases. For example, the A19W mutation onIFNR2 reduced ka by 4-fold, a reduction that clearly relatesto conformational rearrangement during the process ofassociation (as verified using double-mutant cycle analysiswith the W100A mutation on IFNAR2).144 Another examplefor conformational rearrangement rather than electrostaticsaffecting association rates upon mutation was shown forMICA (a major histocompatibility complex-like protein)undergoing a structural transition from disorder to order uponbinding its immuno-receptor, NKG2D.145 In contrast, the rate-limiting step for the human transcription factor pKID domainof CREB, which also undergoes a folding transition uponbinding to the KIC domain of the coactivator CBP, appearsto be formation of a transient complex.146,147 Stabilizing thestructure of the pKID domain prior to binding does notincrease the association rate constant. Also, for complexformation between fas and AChE, molecular dynamicssimulations suggested that fas undergoes a conformationalrearrangement during the binding to AChE, but the rear-rangement does not appear to slow down the associationrate.148,149

    4.2.2. Proteomics View of ‘Hotspots’ for Association

    As discussed above, the association rate constant of aprotein complex depends on a basal rate constant and themagnitude of electrostatic attraction, which can be calculatedin a simplified fashion from the structure of the nativecomplex. In an analysis on a database of 68 heterodimericcomplexes using HyPare it was found that in about one-halfof the complexes electrostatic contributions to the associationrate constants were small (100-fold increases). Defining a residue as ahotspot for association if it changes ka by over 10-fold leavesabout one-half of the complexes without any potential hotspotand a few hotspots per complex in the other complexes. Ofthese putative hotspot residues, about 40% are calculated toincrease the rate of association upon mutation and thusincrease binding affinity. These data suggest that a majorityof protein-protein complexes are not optimized for fastassociation. Moreover, about 40% of the hotspots forassociation are located outside the binding interfaces, makingthem ideal candidates for protein engineering to achievefaster and tighter binding protein complexes. Hotspot residuesare not evenly distributed among the 20 types of amino acids.About 75% of all hotspots are charged residues. This isunderstandable as a charge-reversal mutant changes the totalcharge by two. More intriguing is the small number ofhydrophobic residues that are hotspots, in comparison topolar residues. For 18 out of the 68 complexes in the databaseexperimental values of ka were available, from which basalrate constants were calculated to be in the range from 104 to107 M-1 s-1. Some of these basal rate constants werecorroborated by both salt and mutant data (see above). The

    852 Chemical Reviews, 2009, Vol. 109, No. 3 Schreiber et al.

  • basal rate constants seem to correlate with the sizes of theproteins and shapes of the interfaces (see the next section).

    4.3. Prediction of Electrostatic Rate Enhancementby the Transient-Complex Theory

    Let us now discuss the rigorous prediction of the associa-tion rate constant from the transient-complex theory. Thetheory is given by eq 20 along with a structural model forthe ensemble of the transient complex, which is describedin section 3.1. Transient-complex theory offers two importantadvantages. First, unlike in previous theoretical approachesthe model for the transient complex is based on rigoroustheoretical considerations and is uniquely defined. Second,electrostatic interactions between two associating proteinsare treated directly by the Poisson-Boltzmann equationwithout approximations.

    Transient-complex theory was put to a comprehensive testagainst experimental data4-6,128 for the association rateconstants of four protein pairs (shown in Figure 1) and 23of their mutants over wide ranges of ionic strength.13 Foreach protein pair 100 configurations were randomly selectedfrom the transient-complex ensemble to calculate the averageelectrostatic interaction energy, 〈Uel〉*. Equation 20 was thenused to predict the association rate constant. As shown inFigure 9a the ionic strength dependences of the association

    rates for all four protein pairs are predicted well by thetransient-complex theory. Moreover, the predictions for 23mutants at various ionic strengths agree closely withexperimental results (Figure 9b). In all there are 81 datapoints in the latter comparison, spanning 4 orders ofmagnitude in association rate. The theory thus appears tofulfill the promise of having truly predictive power. It revealsthat among the protein pairs and their mutants studied thebasal rate ka0 can differ by ∼20-fold, but the bulk of thevariations in ka is due to the variations in 〈Uel〉*, which rangesfrom 0 to -6 kcal/mol (the last value translates into a 104-fold rate enhancement).

    In the rigorous calculation of ka it was found that the samebasal rate constant, 1.4 × 106 M-1 s-1, explains both saltand mutation data on barnase-barstar association. Therelatively high basal rate constant for the barnase-barstarcomplex compared to other protein pairs is reproduced byBrownian dynamics simulations.1,13 The variations of basalrate constants among protein complexes can be explainedby the structures of the transient complexes. A higher ka0value, as found for barnase-barstar association, correspondsto a more open ensemble of the transient complex.1

    The question of whether the predicted ka is sensitive tothe precise specification of the transient complex was alsoaddressed.13 As explained in section 3.1, the transientcomplex is specified by the contact level, Nc*, at the onsetof a sharp increase in transltional/rotational freedom (Figure3b). For the barnase-barstar complex the value of Nc* wasdetermined to be 14. The value of Nc* was artificially variedfrom 10 to 18, and at each value ka0 and 〈Uel〉* werecalculated. When combined according to eq 20, the predictedka was found to vary little (

  • in question is the definition of the boundary between theprotein low dielectric and solvent high dielectric. In manyPoisson-Boltzmann calculations this dielectric boundary isspecified by the molecular surface. An alternative choice isthe van der Waals surface.154,155 The difference between thetwo surfaces lies in the many crevices accessible to aspherical solvent probe, which are treated as part of the solutedielectric in the molecular surface protocol but as part ofthe solvent dielectric in the van der Waals surface protocol.It is known that the sign of the electrostatic interaction energyof two oppositely charged subunits can be reversed betweenthe two protocols, positive in the former and negative in thelatter.110,156-158

    The results shown earlier for the average electrostaticinteraction energies in the transient complexes were allobtained with the van der Waals surface protocol. With themolecular surface protocol the same sign reversal mentionedabove was seen on 〈Uel〉*.1,26 Hence, electrostatic rateenhancement now turned into rate retardation. For example,for the barnase-barstar pair when the ionic strength wasvaried from 13 to 2000 mM 〈Uel〉* calculated with the van

    der Waals surface protocol varied from -3.30 to -0.82 kcal/mol. Correspondingly, 〈Uel〉* calculated with the molecularsurface protocol varied from 2.50 to 5.13 kcal/mol. For thelatter results to be consistent with the experimental data forthe association rate constant would require a basal rateconstant in the order of 1010-1011 M-1s-1, which is clearlyunrealistic.

    The transient-complex theory lays out a framework forunderstanding the wide variation in association rate constants(Figure 1) and elucidating the pathways of protein associationand dissociation. As documented above, the theory rational-izes the insensitivity of kd to ionic strength and has achievedgreat success in making quantitative predictions of ka forassociation reactions where conformational rearrangementsare not rate limiting. However, quantitative calculations forrate constants that involve formation or breakup of short-range stereospecific interactions are beyond the scope of thetransient-complex theory.

    5. Interaction Dynamics of Membrane-AnchoredProteins

    Key processes in the initiation of transmembrane signalinginvolve ligand-induced interactions between receptor pro-teins, which are anchored to the membrane. Compared tointeraction kinetics in solution, anchoring of proteins intothe membrane dramatically changes the mobility of theinteraction partners both in terms of diffusion constants(which are reduced by 2 orders of magnitude) and inrestricting the movements to two dimensions (thus actuallymaking the mutual search problem simpler). In general,reduction in dimensionality as well as potentially favorablepreorientation has been proposed to explain the rapidinitiation of signaling.37,159-161 However, few quantitative

    Figure 10. U1 protein-RNA system. (a) The native complex. (b)Representative configurations in the transient complex. The U1Aprotein is represented by the electrostatic surface. Notice that theRNA is moved away from the protein in the transient complex.(Reprinted with permission from ref 26. Copyright 2008 AmericanChemical Society.)

    Figure 11. Comparison of predicted and experimental results forthe binding rates of the wild-type U1 system and eight of itsmutants. Locations of the mutated residues are shown in the inset.(Reprinted with permission from ref 26. Copyright 2008 AmericanChemical Society.)

    854 Chemical Reviews, 2009, Vol. 109, No. 3 Schreiber et al.

  • experimental studies of 2-dimensional protein interactionshave been reported. For IL-4-induced cross-linking of theIL-4 RR with the common γ chain on live cells a surprisinglylow efficiency of association of the low-affinity γ-chainsubunit with the highly stable complex of IL-4 and IL-4 RRon the membrane was observed.162 Recent systematic studiesof ternary cytokine-receptor complexes tethered on artificialmembranes have provided some quantitative understandingof 2-dimensional interactions. These data confirmed theregulatory role of binding affinities for complex formationon membranes. Equilibrium dissociation constants on theorder of 1-10 µM for the interaction in solution yielded2-dimensional equilibrium dissociation constants on the orderof 10-100 molecules/µm2.163 Kinetic analysis of suchcomplexes revealed that the probability of successful colli-sions is indeed higher on membranes, which can be ascribedto the preorientation and longer lifetime of the transientcomplex. This effect is, however, largely compensated bythe decreased diffusion rate, which reduces the number ofcollisions on the same order of magnitude. The orientationof the proteins significantly affects the rate constants ofassociation,164 though protein domains interacting outside themembrane are often rather flexible. Overall, the associationrates at a given average distance of the molecules in solutionand on the surface are surprisingly similar.165 In the dis-sociation process separation of the interacting species bydiffusion plays a critical role. This agrees with the notionthat the association kinetics is highly diffusion controlled.As a consequence, local membrane fluidity in membranemicrodomains may play a critical role for stabilizingmembrane protein complexes. Thus, the energy landscapesof cellular membrane protein complexes are highly influencedby the microcompartmentalization of the membranes. Crowd-ing by by-stander proteins in the membranes may also havea significant effect (cf. section 6).

    Single-molecule fluorescence techniques may contributeto the study of association within membranes. The processof dimerization of epidermal growth factor receptors (EGFR)in cellular membranes and binding of their substrate,epidermal growth factor (EGF), was studied recently usingsingle-molecule fluorescence.166 Fluorescent EGF moleculesbound to EGFR molecules were tracked. Association of asecond EGF molecule to a preformed EGF-EGFR complexwas seen as a sudden doubling of the fluorescence intensity.It was concluded that EGFR dimerization preceded bindingof a second EGF molecule to a complex. The kinetics of thedimerization process itself, however, is yet to be studied.

    6. Roles of the Depletion Effect and Crowding inAssociation Kinetics

    Since protein-protein association mostly takes placewithin cells it is important to understand the particularproperties of the cellular environment which may have aprofound effect on such reactions.167 Most importantly, thecytoplasm of cells is a dense solution of macromolecules ofmany sizes and shapes. In fact, macromolecules may occupyas much as 40% of the total intracellular volume.168 These“crowding agents”, as they are sometimes called, includemostly folded proteins and nucleic acids. Structural proteinsform dynamic networks of filaments, creating the cytoskel-eton and thus dividing the cell into discrete bulk areas ofcytoplasm, where soluble proteins can interact. In addition,there are many ‘natively unfolded proteins’ whose random-

    coil-like conformations can add complexity to the cytoplas-mic milieu.169

    There are two major outcomes of the high intracellularconcentration of macromolecules in their various forms. First,the macroscopic viscosity of the cytoplasm is significantlyhigher than the viscosity of a dilute aqueous solution. Thishigher viscosity should affect long-range translational dif-fusion of proteins. On small length scales there should be alarge heterogeneity in the density of macromolecules, leadingto microscopic viscosity, which depends strongly on position.The crowded cytoplasm may also affect rotational motionof proteins, which is also crucial for protein-proteinassociation. Since rotational diffusion is by nature more localthan translational diffusion and occurs on faster time scales,it is likely to be affected differently by the crowdedcytoplasm.170,171 A second effect of the crowded cytoplasmicsolution is depletion, which can be seen as an effectiveosmotic pressure exerted on all diffusing molecules. Thisphenomenon, which is entropic in nature, is due to the factthat protein solutes exclude (or deplete) some of the solutionvolume from other solutes. These effects are the subjects ofthe following sections.

    6.1. Translational and Rotational Diffusion ofProteins in Concentrated Solutions and Cells

    Dense macromolecular solutions should be discussed moreappropriately in terms of their viscoelasticity rather than theirviscosity.172 Viscoeleastic solutions possess time-scale- andlength-scale-dependent viscosity. The effective viscositysensed by a tracer molecule undergoing fast, short-rangemotion is therefore different than that sensed by a particlemoving slowly over large distances. Further, the cytoplasmicenvironment is not homogeneous, and the density of mac-romolecules and other intracellular components may varyfrom one spot to another within the cell. Thus, diffusion timesmay scale with distance in an unexpected way. Recentexperiments have probed the viscoelasticity of cellularinteriors using microrheology techniques (for reviews seerefs 173 and 174). Of most interest to the present revieware experiments that directly measure diffusion coefficientsinside cells or in complex solutions that may mimic thecytoplasm.

    A useful probe for studying translational and rotationaldiffusion inside cells is green fluorescen