Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Feb 1.
Published in final edited form as: Curr Opin Struct Biol. 2016 Oct 15;42:1–5. doi: 10.1016/j.sbi.2016.10.001

Visualizing the nanoscale: Protein internal dynamics and neutron spin echo spectroscopy

David J E Callaway 1,*, Zimei Bu 1,*
PMCID: PMC5374024  NIHMSID: NIHMS823521  PMID: 27756047

Abstract

The most complex molecular machines are proteins found within cells. Protein dynamics, in particular dynamics on nanoscales, presents us with a novel paradigm for cell signaling: the idea that proteins and protein complexes can communicate directly within themselves to effect long-range information transfer, via coupled domains and correlated residue clusters. This idea has been little explored, in large part because of a paucity of experimental techniques that can address the necessary questions. Here we review recent progress in developing a promising new approach, neutron spin echo spectroscopy.

Keywords: protein dynamics, nanoscale protein motion, nonequilibrium statistical mechanics, NHERF1, ezrin, ERM proteins, PDZ domain, multidomain proteins

Graphical abstract

graphic file with name nihms823521u1.jpg

Proteins MOVE!

Far from the static objects seen in textbooks, proteins are dynamic actors whose internal dynamics stimulates and controls a medley of essential biological processes. The natural scale for these internal motions is that of nanoseconds to microseconds and nanometers, so we therefore refer to them as nanoscale motions. Paradoxically, the visualization of this nanoscale activity requires using large-scale sophisticated neutron scattering techniques at spallation neutron sources and nuclear reactors. Our course of action is thus very much an adventure in Big Science, and requires traveling through an interdisciplinary conurbation between nuclear physics, nonequilibrium statistical mechanics, molecular biology and protein chemistry in order to reach its goals. We therefore give an outline of the field in hopes that the reader will share our enthusiasm for this burgeoning endeavor. We begin by summarizing basic properties of nanoscale protein motions.

Protein motions are overdamped, creeping movements rather than underdamped oscillations

The environment in which proteins act is one of low Reynolds number (usually abbreviated Re), which is the ratio of the magnitude of inertial forces to that of the forces that arise from the viscous drag that opposes motion. If inertial forces are more important, Reynolds number is large, and forces are proportional to mass times acceleration. If viscous drag is more important, the Reynolds number is small (< 1000), and mechanical forces are proportional to the velocity of the protein, incorporating a concept known as the mobility tensor. Reynolds number can be estimated by the simple formula Re=LV/ν, where L is a characteristic length scale of the protein, V is a characteristic velocity, and ν is the kinematic viscosity of the solvent (for water, ν = 105 Å2/ns). Simple calculations show that even a complex as large as a ribosome is still in the overdamped regime (1)**. The environment of a protein thus has more in common with playing badminton at the bottom of a swimming pool full of molasses (low Re) than in crossing the Atlantic in the Titanic (high Re).

Proteins obey Brownian dynamics

Protein dynamics arises as a result of an interplay between the mechanical forces mentioned above, and the thermal forces that arise from the collision of the protein with solvent molecules. These thermal forces are random in magnitude and direction, and lead to the protein undergoing diffusion. A freely diffusing object displays what is called Brownian motion, with frequent changes in the direction and speed of its movement. The world in which proteins operate is therefore characterized by the presence of a significant amount of noise and the resultant diffusion of protein subunits arising from thermal motion. This thermal motion is essential for the protein to reach its equilibrium state (2)**.

The relaxation timescales of these overdamped internal diffusive motions are of order L2 /D, where D is the overall diffusion constant of the protein and L is the characteristic length scale of the motion. This follows from a Brownian analysis of simple harmonic motion (3)*. According to the domain concept of structural biology (4) and the references therein, proteins can be considered as being comprised of somewhat rigid domains connected by soft spring linkers. Since protein diffusion constants are of the order of a few Å2/ns, we see that for domains connected by linkers of the order of L ∼ 10 Å or so, the relaxation timescales of nanometer internal modes will be of the order of nanoseconds, with rotational relaxation times that are a factor of 4 or so longer. This justifies our use of the word nanoscales for these motions. The scale of such motions is optimal for neutron spin echo spectroscopy (5*,6**,7*).

Neutron spin-echo spectroscopy (NSE) is unique in its capacity to determine nanoscale motions

NSE is a quasielastic neutron scattering technique that employs the Larmor precession of neutron spins in a magnetic guide field as a clock to measure extremely small changes in velocities of scattering neutrons (8,9), and thus enables the detection of very small energy changes corresponding to nanosecond-to-microsecond dynamics. This is a dynamic regime that is difficult or impossible to observe with other methods (10).

Neutron spin echo spectroscopy measures the intermediate scattering function I(Q,t), which is the spatial Fourier transformation of the space-time van Hove correlation function G(r,t)(2), I(Q,t)=VG(r,t)exp(iQr)dr, where Q is the scattering vector. (The designation “intermediate” arises precisely because only one of the variables of G(r,t) is Fourier transformed).

Nanoscale motions determine I(Q,t)

For a given Q, I(Q,t) typically can be fit to a single exponential in time (and is difficult to fit to more exponentials, because of the problem of separating rotational diffusion from internal mode relaxation). A natural way to interpret neutron scattering data is therefore to examine the effective diffusion constant Deff(Q) as a function of Q, which is determined by the normalized intermediate scattering function I(Q,t)/I(Q,0):

Γ(Q)=limt0tln[I(Q,t)/I(Q,0)]Deff(Q)=Γ(Q)Q2 Eq. 1

where I(Q,0) is the static form factor. As I(Q,t)/I(Q,0) is generally amenable to a single-exponential fit in time (see Figure 1), Deff(Q) can be accurately estimated by the first cumulant expression (5,6,11,12)*:

Figure 1. NHERF1 alone can be described by a rigid-body model.

Figure 1

Comparing experimental Deff(Q) of NHERF1 (black open squares) with rigid-body calculation (black solid line).

Deff(Q)=kBTQ2jlbjbl(QHjlTQ+LjHjlRLl)eiQ(rjrl)jlbjbleiQ(rjrl) Eq. 2

which is a generalization of the remarkable Akcasu-Gurol (AG) formula (2,13*,14*) to rotational motion (5). (This formula is an elementary sum-rule result that results from the fact that the associated Smoluchowski equation forms a Sturm-Liouville system). Here, bj is the scattering length of a subunit j, HT is the translational mobility tensor, and HR is the rotational mobility tensor. The coordinates of the various subunits (“subunits” are scattering centers that can be atoms, beads, or domains, generically called beads here) are taken relative to the center of friction of the protein, and are given by rj (note that Σ rj = 0); kBT is the usual temperature factor; and Lj= rj × Q is the torque vector for each coordinate. The brackets < > denote an equilibrium average over all protein conformations and scattering lengths, as well as an orientational average over the vector Q, so that <Qa Qb exp (i Qr)> Q-2 = (1/3)δabj0(Qr)+[(1/3)δab - (ra rb /r2)] j2(Qr) can be expressed in terms of spherical Bessel functions j.

The AG approach described in Eq. 2 is valid for either rigid bodies or rigid-body subunits connected by soft spring linkers (5,11). The translational mobility tensor HT is defined by the velocity response v = HT F to an applied force F. The rotational mobility tensor HR is defined by the angular velocity response ω = HR τ to an applied torque τ. In practice, the structural coordinates of a protein may be obtained from high-resolution crystallography or NMR or from low-resolution EM, SAXS or SANS. Comparison of the calculations Eq. 2 to experimental Deff(Q) thus allows one to test models of the mobility tensors. The rotational mobility tensor HR can be determined from the translational mobility tensor HT (6,15). We point out that the first cumulant expression is explicitly independent of the spring constant of a linker connecting the domains. It is thus not necessary to fit multiple time exponentials to separate translational, rotational, and internal modes in order to reveal internal dynamics.

The mobility tensor directly reveals internal degrees of freedom

For a rigid body composed of N identical subunits, the translational mobility tensor HT is a matrix with N2 identical 3×3 rotation elements. This must be so, since HT yields the velocity response of e.g., subunits B and C to a force applied to subunit A. If the mobility tensor components HAB and HAC are unequal, the velocity response of B and C will be different, B and C will move apart, and the body will no longer remain rigid. Additionally, we showed that for a rigid, uniform (nondeuterated) body, Deff(Q→∞) = 2 Deff(Q=0)(6,12). Large scale numerical simulations are thus not required to demonstrate internal dynamics.

Comparing the calculated Deff(Q) with data allows one to extract the relative degree of dynamic coupling between the various components of the system, for this dynamic coupling is defined by the mobility tensor. For example, a rigid two-domain system has a translational mobility tensor

H=H0(1111) Eq. 3a

with all elements of the tensor equal, and yields [via Eq. 2] the simple result that the translational contribution to the effective diffusion constant is given by DTeff(Q)= kBT H0, independent of Q. By contrast, a two-domain flexible protein with internal motion has a translational mobility tensor

H=(H100H2) Eq. 3b

in principal coordinates. The application of equal forces to the two domains of a flexible protein will then result in their having different velocities, revealing internal motion. For the case where there is one internal translational mode between subunits 1 and 2 with D1=kBTH1 and D2= kBTH2 (5), the translational contribution to the effective diffusion constant for this flexible system is:

DTeff(Q)=D1S1(Q)+D2S2(Q)S(Q) Eq. 3c

Here, S1(Q) and S2(Q) are the form factors of the separate individual protein domains, while S(Q) is the form factor of the entire protein. Orientational averages are performed, so that, e.g., S(Q)=Σj1 j0 (Qrj1); and S(Q=0)=N2 and S(Q→∞) = N, with N the number of beads. Since DTeff(Q→∞) = DReff(Q→∞) (DReff(Q) is the rotational contribution to Deff(Q)) while DReff(Q=0) = 0, Deff(Q→∞) = 4 Deff(Q=0) (6).

If one domain is a fraction × of the whole protein then Eq (2) yields

Deff(Q)/Deff(Q=0)2[x2+(1x)2]1 Eq. 4

Similarly the parsing of a polysyndetonic uniform protein into M equal domains gives Deff(Q→∞) ∼ 2M Deff(Q=0), related to the transition from coherent to incoherent scattering. The effect can be amplified by selective deuteration (6,16). The Q dependence of the effective diffusion constant is thus a highly sensitive probe of active internal dynamic modes and correlations (12,17).

The 3×3 matrix HR is evaluated here by an inversion of a 3×3 matrix calculated by summing over N bead coordinates n (6) A compact formula arises (6) in terms of a 3×3 matrix inverse, using the protein diffusion constant D0 and the N structural coordinates of the protein (defined so that Σ n rn = 0):

HRαβ=N(D0/kBT)[n(δαβr2nrrαrnβ)]1 Eq. 5

NHERF1: A paradigm of long-range allostery

The multi-PDZ domain scaffolding NHERF1 interacts with a number of physiologically important transmembrane receptors and ion transport proteins in epithelial cells (18,19). NHERF1 binds to the membrane-cytoskeleton linker protein Ezrin. Previously, we uncovered a prominent long-range allosteric regulatory behavior in NHERF1: Upon Ezrin binding to NHERF1, the binding affinities of the PDZ domains of NHERF1 for target membrane proteins are dramatically increased, even though the PDZ domains are located more than 120 Å away from the Ezrin-binding site (6,20,21*). Thus, Ezrin allosterically modulates NHERF1 to assemble membrane protein complexes and to tether the assembled complexes to the cytoskeletal actin network, which modulates the cell surface targeting and intracellular trafficking of the assembled signaling complexes (22-24).

Long-range allostery is an important consequence of internal protein motion, and can be revealed by NSE

Cell signaling ultimately involves cytoskeletal rearrangement via structure sensing and a Rube Goldberg conglomeration of gear selectors, rheostats, and leaky valve faucets. Scaffolding and adapter proteins not only provide a scaffold to dock signaling partners, but also propagate and relay signals allosterically to the site of regulation over a long distance. Our recent study using neutron spin echo spectroscopy reveals that inter-domain motions in NHERF1 on nanometer length-scales and on submicrosecond time scale can propagate allosteric binding signals dynamically. The long-range allostery and nanoscale protein domain dynamics during the interactions of NHERF1 and ezrin provide a paradigm for the mechanisms of how cellular signals can be transmitted in a cellular signaling network (6,12).

NSE reveals the activation of NHERF PDZ domain motion by FERM domain binding

First, we present the results for a rigid body, see Figure 1. We then present the calculation of the model representing the deuterated NHERF1·dFERM and hydrogenated NHERF1·hFERM complexes. Details are given in ref. (6). Figure 2A compares the NSE data with the Deff(Q) from the rigid body model for the hydrogenated and partially deuterated complexes. Figure 2B is the Deff(Q) of the model incorporating internal domain motion between PDZ1 and the rest of the complex. This shows that the long-range allosteric motion is activated by FERM binding.

Figure 2. Selective deuteration highlights the activation of domain motion in NHERF1 upon binding to Ezrin FERM domain.

Figure 2

(A) Comparing NSE data with rigid model calculations for the NHERF1·dFERM and NHERF1·hFERM complexes using the coordinates of the docked domains. (B) Comparing experimental Deff(Q) with calculations incorporating interdomain motion (via the mobility tensor) between PDZ1 and PDZ2, for NHERF1·dFERM (dashed red line) and NHERF1·hFERM (dashed blue line).

Deuterium labeling of a domain amplifies the effects of internal motion detected by NSE by masking a portion of the form factor, and, as a result, highlighting the contribution of the terms of internal domain motion (6,15). Because for proteins inertial forces are less important than diffusive, viscous effects, protein dynamics should be largely independent of the mass of the protein. The diffusion constant of a deuterated protein is the same as a hydrogenated protein, even though deuterium has twice the mass of hydrogen. Deuterium contrast matching is thus valuable in neutron spin echo spectroscopy experiments.

New horizons: The highly dynamical nature of protein complexes

Protein dynamics presents us with the compelling idea that proteins can communicate within themselves to effect long-range allosteric information transfer (6,25). Experimental techniques to observe these essential phenomena in proteins macromolecular complexes are however in their infancy (26-28). This review discusses an especially promising candidate: neutron spin-echo spectroscopy. We believe that it is impossible not to be excited about the challenges ahead, and the rewards for their successful solution.

Highlight.

  • Nanoscale protein dynamics affects cell signaling propagation via long-range allostery in a largely unknown fashion.

  • It is difficult at best to probe protein dynamics at these nanoscales since techniques are lacking.

  • The novel technique of neutron spin echo spectroscopy (NSE) reveals nanoscale dynamics changes that control the assembly of protein complexes.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1**.Howard J. Mechanics of Motor Proteins and the Cytoskeleton. Sinauer Associates; Sunderland, MA: 2001. Large macromolecular machines utilize overdamped motions. See chapter 3, especially Table 3.4 on page 39. [Google Scholar]
  • 2**.Doi M, Edwards SF. The theory of polymer dynamics. Oxford University Press; Oxford; 1986. pp, Excellent introduction to nonequilibrium statistical mechanics of macromolecular dynamics. [Google Scholar]
  • 3*.Wax N. In: (1954) Noise and stochastic processes. Wax Nelson., editor. Vol. 1. New York: Dover Publication; 1954. Classic text on Brownian motion. [Google Scholar]
  • 4.Dziubinski M, Daniluk P, Lesyng B. ResiCon: a method for the identification of dynamic domains, hinges and interfacial regions in proteins. Bioinformatics. 2016;32:25–34. doi: 10.1093/bioinformatics/btv525. [DOI] [PubMed] [Google Scholar]
  • 5*.Bu Z, Biehl R, Monkenbusch M, Richter D, Callaway DJ. Coupled protein domain motion in Taq polymerase revealed by neutron spin-echo spectroscopy. Proc Natl Acad Sci U S A. 2005;102:17646–17651. doi: 10.1073/pnas.0503388102. Introduces the mobility tensor into NSE analysis. The mobility tensor determines the dynamical coupling between protein domains. Taq polymerase utilizes correlated domain dynamics over 70 Å to coordinate nucleotide synthesis and cleavage during DNA synthesis and repair. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6**.Farago B, Li J, Cornilescu G, Callaway DJ, Bu Z. Activation of nanoscale allosteric protein domain motion revealed by neutron spin echo spectroscopy. Biophys J. 2010;99:3473–3482. doi: 10.1016/j.bpj.2010.09.058. Upon binding to a partner, nanoscale domain motion between PDZ domains is allosterically activated. Selective deuteration of a subunit can highlight and amplify internal motion for NSE detection. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7*.Callaway DJ, Farago B, Bu Z. Nanoscale protein dynamics: A new frontier for neutron spin echo spectroscopy. The European physical journal E, Soft matter. 2013;36:9891. doi: 10.1140/epje/i2013-13076-1. Emphasizes the high potential of NSE. [DOI] [PubMed] [Google Scholar]
  • 8.Mezei F. Neutron Spin Echo: proceedings of a Laue-Langevin Institut workshop. Springer; Heidelberg: 1980. The Principles of Neutron Spin Echo. [Google Scholar]
  • 9.Farago B. Time-of Flight Neutron Spin Echo: Present Status. In: Mezei F, Pappas C, Gutberlet T, editors. Neutron Spin Echo Spectroscopy. Springer; 2003. pp. 15–34. [Google Scholar]
  • 10.Berger CL. Breaking the millisecond barrier: single molecule motors wobble to find their next binding sites. Biophys J. 2013;104:1219–1220. doi: 10.1016/j.bpj.2013.02.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Callaway DJ, Bu Z. Nanoscale protein domain motion and long-range allostery in signaling proteins- a view from neutron spin echo sprectroscopy. Biophys Rev. 2015;7:165–174. doi: 10.1007/s12551-015-0162-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Bu Z, Callaway DJ. Proteins MOVE! Protein dynamics and long-range allostery in cell signaling. Adv Protein Chem Struct Biol. 2011;83:163–221. doi: 10.1016/B978-0-12-381262-9.00005-7. [DOI] [PubMed] [Google Scholar]
  • 13*.Akcasu Z, Gurol H. Quasi-Elastic Scattering by Dilute Polymer-Solutions. Journal of Polymer Science Part B-Polymer Physics. 1976;14:1–10. Derivation of AG first cumulant formula. [Google Scholar]
  • 14*.Fixman M. Variational bounds for polymer transport coefficients. J Chem Phys. 1983;78:1588. Highly readable explication of implications of Smoluchowski equation, such as AG result. [Google Scholar]
  • 15.Callaway DJ, Bu Z. Essential Strategies for Revealing Nanoscale Protein Dynamics by Neutron Spin Echo Spectroscopy. Methods Enzymol. 2016;566:253–270. doi: 10.1016/bs.mie.2015.05.011. [DOI] [PubMed] [Google Scholar]
  • 16.Callaway DJ, Bu Z. Essential Strategies for Revealing Nanoscale Protein Dynamics by Neutron Spin Echo Spectroscopy. Methods in Enzymology. 2015 doi: 10.1016/bs.mie.2015.05.011. [DOI] [PubMed] [Google Scholar]
  • 17.Bu Z, Cook J, Callaway DJ. Dynamic regimes and correlated structural dynamics in native and denatured alpha-lactalbumin. J Mol Biol. 2001;312:865–873. doi: 10.1006/jmbi.2001.5006. [DOI] [PubMed] [Google Scholar]
  • 18.Weinman EJ, Hall RA, Friedman PA, Liu-Chen LY, Shenolikar S. The association of NHERF adaptor proteins with g protein-coupled receptors and receptor tyrosine kinases. Annu Rev Physiol. 2006;68:491–505. doi: 10.1146/annurev.physiol.68.040104.131050. [DOI] [PubMed] [Google Scholar]
  • 19.Fehon RG, McClatchey AI, Bretscher A. Organizing the cell cortex: the role of ERM proteins. Nat Rev Mol Cell Biol. 2010;11:276–287. doi: 10.1038/nrm2866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Li J, Dai Z, Jana D, Callaway DJ, Bu Z. Ezrin controls the macromolecular complexes formed between an adapter protein Na+/H+ exchanger regulatory factor and the cystic fibrosis transmembrane conductance regulator. J Biol Chem. 2005;280:37634–37643. doi: 10.1074/jbc.M502305200. [DOI] [PubMed] [Google Scholar]
  • 21*.Li J, Callaway DJ, Bu Z. Ezrin induces long-range interdomain allosteryin the scaffolding protein NHERF1. J Mol Biol. 2009;392:166–180. doi: 10.1016/j.jmb.2009.07.005. Biochemical evidence supporting NSE results. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Loureiro CA, Matos AM, Dias-Alves Â, Pereira JF, Uliyakina I, Barros P, Amaral MD, Matos P. A molecular switch in the scaffold NHERF1 enables misfolded CFTR to evade the peripheral quality control checkpoint. Sci Signal. 2015;8:ra48–ra48. doi: 10.1126/scisignal.aaa1580. [DOI] [PubMed] [Google Scholar]
  • 23.Abbattiscianni AC, Favia M, Mancini MT, Cardone RA, Guerra L, Monterisi S, Castellani S, Laselva O, Di Sole F, Conese M. Correctors of mutant CFTR enhance subcortical cAMP–PKA signaling through modulating ezrin phosphorylation and cytoskeleton organization. J Cell Sci. 2016;129:1128–1140. doi: 10.1242/jcs.177907. [DOI] [PubMed] [Google Scholar]
  • 24.Lobo MJ, Amaral MD, Zaccolo M, Farinha CM. EPAC1 activation by cAMP stabilizes CFTR at the membrane by promoting its interaction with NHERF1. J Cell Sci, jcs. 2016 doi: 10.1242/jcs.185629. 185629. [DOI] [PubMed] [Google Scholar]
  • 25.Cooper A, Dryden DTF. Allostery without conformational change, a plausible model. Eur Biophys J. 1984;11:103–109. doi: 10.1007/BF00276625. [DOI] [PubMed] [Google Scholar]
  • 26.Hong L, Sharp MA, Poblete S, Biehl R, Zamponi M, Szekely N, Appavou MS, Winkler RG, Nauss RE, Johs A, Parks JM, Yi Z, Cheng X, Liang L, Ohl M, Miller SM, Richter D, Gompper G, Smith JC. Structure and dynamics of a compact state of a multidomain protein, the mercuric ion reductase. Biophys J. 2014;107:393–400. doi: 10.1016/j.bpj.2014.06.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Smolin N, Biehl R, Kneller GR, Richter D, Smith JC. Functional domain motions in proteins on the ∼1-100 ns timescale: comparison of neutron spin-echo spectroscopy of phosphoglycerate kinase with molecular-dynamics simulation. Biophys J. 2012;102:1108–1117. doi: 10.1016/j.bpj.2012.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Lal J, Fouquet P, Maccarini M, Makowski L. Neutron spin-echo studies of hemoglobin and myoglobin: multiscale internal dynamics. J Mol Biol. 2010;397:423–435. doi: 10.1016/j.jmb.2010.01.029. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES