Biomolecular hydrodynamics

Similar documents
Biomolecular hydrodynamics

Biomolecular hydrodynamics

Hydrodynamics: Viscosity and Diffusion Hydrodynamics is the study of mechanics in a liquid, where the frictional drag of the liquid cannot be ignored

Analytical Ultracentrifugation. by: Andrew Rouff and Andrew Gioe

Measuring S using an analytical ultracentrifuge. Moving boundary

Transport Properties: Momentum Transport, Viscosity

By: Ashley and Christine Phy 200 Professor Newman 4/13/12

Use of SEC-MALS. (Size Exclusion Chromatography - Multi Angle. Light Scattering) for protein quality and characterization

Solution structure and dynamics of biopolymers

SPRING 2011 Phys 450 Solution set 2

How DLS Works: Interference of Light

Measuring the size and shape of macromolecules. Hydrodynamics: study of the objects in water How do the move? Translation Rotation

8.6 Drag Forces in Fluids

C C C C 2 C 2 C 2 C + u + v + (w + w P ) = D t x y z X. (1a) y 2 + D Z. z 2

Viscometry. - neglect Brownian motion. CHEM 305

Chem 406 Biophysical Chemistry Lecture 1 Transport Processes, Sedimentation & Diffusion

Chem/Biochem 471 Exam 2 11/14/07 Page 1 of 7 Name:

Lecture 6 Molecular motion and Transport p roperties properties

Part I.

University of Washington Department of Chemistry Chemistry 453 Winter Quarter 2008

UltraScan Workshop/Bioc5083 Hydrodynamic Methods

What s important: viscosity Poiseuille's law Stokes' law Demo: dissipation in flow through a tube

There are 6 parts to this exam. Parts 1-3 deal with material covered since the midterm exam. Part 4-6 cover all course material.

BAE 820 Physical Principles of Environmental Systems

UNIVERSITY OF OSLO FACULTY OF MATHEMATICS AND NATURAL SCIENCES

Equilibrium Sedimentation & Sedimentation Velocity: Random Walks in the presence of forces.

MAE 545: Lecture 2 (9/22) E. coli chemotaxis

BME Engineering Molecular Cell Biology. Basics of the Diffusion Theory. The Cytoskeleton (I)

CHAPTER V. Brownian motion. V.1 Langevin dynamics

SEC MALLs and AUC. 2. Conformation and flexibility Viscometry,

Diffusion-deconvoluted sedimentation coefficient distributions for the analysis of interacting and non-interacting protein mixtures

University of Washington Department of Chemistry Chemistry 355 Spring Quarter 2005

t = no of steps of length s

7 The Navier-Stokes Equations

Brownian motion and the Central Limit Theorem

Introduction to First Order Equations Sections

SIMPLE HARMONIC MOTION

Chem728 Spr. 12 page 1 2/29/12. Assignment 2 Estimates of Hydrodynamic Radii from Dynamic Light Scattering Data. ANSWERS

EXPERIMENT 17. To Determine Avogadro s Number by Observations on Brownian Motion. Introduction

Problem Set Number 01, MIT (Winter-Spring 2018)

Diffusion. Spring Quarter 2004 Instructor: Richard Roberts. Reading Assignment: Ch 6: Tinoco; Ch 16: Levine; Ch 15: Eisenberg&Crothers

2. Molecules in Motion

Forces and Motion in One Dimension

Thursday, August 4, 2011

Lecture 26: More on Gel Filtration Chromatography and the Trypsin Resurrection Experiment

Diffusion in biological systems and Life at low Reynolds number

Diffusion in the cell

The dynamics of small particles whose size is roughly 1 µmt or. smaller, in a fluid at room temperature, is extremely erratic, and is

Introduction To Materials Science FOR ENGINEERS, Ch. 5. Diffusion. MSE 201 Callister Chapter 5

2.3 Damping, phases and all that

MOTION OF BODIES IN FLUIDS

Transport (kinetic) phenomena: diffusion, electric conductivity, viscosity, heat conduction...

HSC PHYSICS ONLINE B F BA. repulsion between two negatively charged objects. attraction between a negative charge and a positive charge

Dr. Christoph Johann Wyatt Technology Europe GmbH Copyright Wyatt Technology Europe GmbH All Rights reserved 1

Brownian Motion and The Atomic Theory

Class 11 Physics NCERT Exemplar Solutions Motion in a Straight Line

Are DNA Transcription Factor Proteins Maxwellian Demons?

2. FLUID-FLOW EQUATIONS SPRING 2019

Chapter 5. Methods for the Separation and Characterization of Macromolecules

Ionization Detectors. Mostly Gaseous Detectors

2 Structure. 2.1 Coulomb interactions

Time-Dependent Statistical Mechanics 1. Introduction

36. TURBULENCE. Patriotism is the last refuge of a scoundrel. - Samuel Johnson

Game Physics. Game and Media Technology Master Program - Utrecht University. Dr. Nicolas Pronost

Phys 450 Spring 2011 Solution set 6. A bimolecular reaction in which A and B combine to form the product P may be written as:

Entropy generation and transport

Part III. Polymer Dynamics molecular models

BROWNIAN DYNAMICS SIMULATIONS WITH HYDRODYNAMICS. Abstract

by Martin Bier, September 2003 past decade microcapillaries have found wide application in analytical

You know for EE 303 that electrical speed for a generator equals the mechanical speed times the number of poles, per eq. (1).

Modeling of a centrifuge experiment: the Lamm equations

Physical Pharmacy. Diffusion

Particles, drops, and bubbles. Lecture 3

Added topics: depend on time/rate - Skipped

University of Washington Department of Chemistry Chemistry 453 Winter Quarter 2013

Particle Characterization Laboratories, Inc.

Transport Properties

SEDIMENTATION VELOCITY ANALYTICAL ULTRACENTRIFUGATION

University of Washington Department of Chemistry Chemistry 453 Winter Quarter 2005

For one such collision, a new particle, "k", is formed of volume v k = v i + v j. The rate of formation of "k" particles is, N ij

Part II => PROTEINS and ENZYMES. 2.7 Enzyme Kinetics 2.7a Chemical Kinetics 2.7b Enzyme Inhibition

Lecture 2. Lecture 1. Forces on a rotating planet. We will describe the atmosphere and ocean in terms of their:

Wiley Plus Reminder! Assignment 1

Rotational Brownian motion; Fluorescence correlation spectroscpy; Photobleaching and FRET. David A. Case Rutgers, Spring 2009

This is a Gaussian probability centered around m = 0 (the most probable and mean position is the origin) and the mean square displacement m 2 = n,or

FORMULA SHEET. General formulas:

Protein Dynamics, Allostery and Function

From a Mesoscopic to a Macroscopic Description of Fluid-Particle Interaction

You should be able to demonstrate and show your understanding of:

CPS Instruments Europe P.O. Box 180, NL-4900 AD Oosterhout, The Netherlands T: +31 (0) F: +31 (0) E:

File ISM02. Dynamics of Soft Matter

PHYS-102 LAB 2. Millikan Oil Drop Experiment

Polymerization and force generation

Math 308 Exam I Practice Problems

Use separation of variables to solve the following differential equations with given initial conditions. y 1 1 y ). y(y 1) = 1

THE SOLUTION CONFORMATION OF NOVEL ANTIBODY FRAGMENTS STUDIED USING THE PROTEOMELAB XL-A ANALYTICAL ULTRACENTRIFUGE

Phys /7/13. November 7, 2013 Physics 131 Prof. E. F. Redish. Theme Music: Duke Ellington Take the A Train. Cartoon: Bill Amend FoxTrot

Physics 3 Summer 1990 Lab 7 - Hydrodynamics

1. Droplet Growth by Condensation

PY1008 / PY1009 Physics Gravity I

Transcription:

Biomolecular hydrodynamics David Case Rutgers, Spring, 2009 February 1, 2009

Why study hydrodynamical methods? The methods we study in this section are low resolution ones one studies diffusional motion and friction to learn about: 1 molecular mass (and hence, aggregation) 2 molecular charge (for separation, or as a function of ph) 3 broad features of molecular shape (especially useful for very anisotropic systems, like DNA, collagen, filaments, etc.) 4 rigidity and flexibility (again, at low resolution) There are two main classes of these methods: driven methods, that apply an external force (gravity, electric field) to the system, and usually come to a steady-state environment, and diffusion or Brownian motion methods, that measure the fluctuations that naturally occur (by thermal motion) in an unforced system at equilibrium.

Life at low Reynolds number Even macro molecules are really quite small, and would ordinarily move very quickly at room temperature, if the were not constantly being buffetted about by the aqueous environment. Because of the viscosity of the environment, inertial forces are not very important in their motion: as Purcell says (p. 256 of your text): [at very low Reynolds number] what You are doing at the moment is entirely determined by the forces that are exerted on You at the moment, and by nothing in the past relative effect of viscosity is represented in the Reynolds number R = ρul/η. For example, for a virus of diameter 5 x 10 6 cm, in water (ρ=1 gm cm 3 ), with visosity η= 0.01 poise = 10 2 gm cm 1 s 1, you get a Reynolds number of 5 x 10 7. (See other examples on p. 256.) The

Frictional coefficients Consider a particle moving with velocity v under the influence of some external force F (say a graviational or electrostatic external force). In addition to the external force, there is a viscous drag inhibiting the motion, which is proportional to the velocity: F f v = m(dv/dt) If the velocity at time 0, v 0 is parallel to the applied force F (here assumed to be constant), then the linear differential equation is easy to solve: v(t) = (F/f ) + [v 0 (F/f )]e ft/m The velocity decays (quickly, if you plug in numbers) to a final value F/f which is linear in the applied force.

Relation between friction and molecular size and viscosity The frictional drag must depend on particle size, and on the viscosity of the medium (typically water, which has a viscosity η of about 0.01 poise = 0.01 g cm 1 s 1 ). Dimensional analysis can help here: suppose for a sphere of radius r: f η x r y Now f has units of g s 1. The only possible values for x and y are 1, so that f ηr. For a sphere, where the molecules of the solvent stick to the surface, messy algebra can find Stokes law: f sph = 6πηr For reference later on, the corresponding friction that retards rotation of a sphere of volume V = 4πr 3 /3 is: f sph,rot = 6ηV

Size and shape dependence of friction coefficients It is common to compare the frictional coefficients of other simple shapes to that of a sphere of the same volume: Note: macromolecules always show a little more friction than they should, based on their size and shape. This is because (roughly speaking!) they drag along a certain amount of water with them (a hydration layer ). See pp. 257-258 and 327-328 of your text.

Friction coefficients for other shapes

(Ultra)centrifuge: simplest idea The centrifugal force on the particles is ω 2 x per unit mass, where ω is the rotor speed and x is the distance from the center of the rotor. The terminal velocity should equal this force divided by the friction coefficient. It is typical in the literature to report the Svedberg coefficient s = v/(ω 2 x), which has units of s 1. Measurements are reported in Svedberg units, where 1S=10 13 s. In simple terms, you can just measure the velocity, and compare s of the sample to that for standards of known molecular weight. However, realistically, the particles in the sample cell are experiencing both driven motion (because of the rotor speed) and diffusion (natural spreading because of concentration gradients. We need to examine diffusion first, then come back to the centrifuge problem later.

Fick s first and second laws of diffusion (Definition The flux J is the mass transported across a boundary per second, divided by the area of the boundary. In the figure, the mass in the little box is cadx (concentration times volume). Flux J is this mass divided by Adt; hence J = c(dx/dt) = cv, where v is the velocity of the flow.) Consider the tiny volume shown in the figure. The rate of mass transport from left to right through the zone must be proportional to the concentration at the left, c(x), and inversely proportional to the thickness of the zone, dx. Transport in the opposite direction is proportion to c(x + dx)/dx. The net rate is given by Fick s first law of diffusion: J = [Dc(x) Dc(x + dx)]/dx = D( c/ x) t (1)

Applying mass balance Now, consider the change in mass inside the volume. This is just dm/dt = J(x) J(x + dx). In terms of concentrations, this is dc/dt = (1/V )(dm/dt) = (1/dx)(dm/dt), where we have taken A to be a unit area. Combining these: (dc/dt) x = [J(x) J(x + dx)]/dx = ( J/ x) t (2) Combining with Fick s first law, we get Fick s second law of diffusion: (dc/dt) x = ( D( c/dx)/dx) = D ( 2 c/ x 2) t (3)

Solution to the diffusion equations If you start with a spike (Dirac delta function) of material, it will spread by diffusion according to c(x,t) = W 0 exp( x 2 /4Dt)/ 4πDt, where W 0 is the total mass.

Diffusion in the presence of an external force Remember that a low Reynolds number, and accelerations caused by external forces are short-lived, and one quickly comes to a terminal velocity v = F/f, where F is the external force and f is the friction coefficent. In the presence of both an external force and diffusion (Fick s first law): J = D ( c/ x) + cf/f (4) By virtue of Eq. 2, we get the Smoluchowski equation: ( c/ t) = D ( 2 c/ x 2) (F/f )( c/ x) (5) Now, suppose you come to equilibrium (J = 0) where diffusion matches the external force (sedimentation equilibrium, say). Here gravity (or centrifugal force) is pulling particles down, creating a concentration gradient (higher concentration at the bottom). But diffusion works in the opposite direction, trying to equalize concentrations, and hence pulling particles up.

The Einstein-Smoluchowski relation Rearranging Eq. 4 with J = 0 gives D dc c = F f dx Integrate both sides of this equation from x = 0 (top of the beaker) to position x, and set w = Fdx where w is the (reversible) work, minus sign because gravity pulls downward. Then D ln c(x) c(0) = w, or c(x) = c(0) exp( w(x)/fd) f But at equilibrium we must also have the Boltzmann distribution law, so fd = kt, or D = kt /f, which is the Einstein-Smoluchowski equation. It is an example of a fluctuation-dissipation relation, connecting fluctuations (D) with friction, (dissipation, f ).

Sedimentation equilibrium: example Tumor necrosis factor alpha (TNF) was the first known member of a family of signaling molecules involved in inflammation, apoptosis, and many other important functions. A hallmark of this family is that these proteins normally occur as trimers in solution. A potential new member of this family was identified on the basis of sequence homology. However, when it was expressed in E. coli and refolded from inclusion bodies, it appeared to be a monomer based on its elution relative to standards on size-exclusion chromatography (SEC). Did this mean it was not truly a member of this family, or simply that it was not correctly refolded, or was the mass estimate from SEC wrong? The graph to the right shows some sedimentation equilibrium data for this molecule, showing the concentration as a function of position within the cell as monitored by absorbance at 230 nm. Note that the total amount of protein for this experiment was <10 micrograms.

More on TNF This next graph shows that data re-plotted as the natural log of absorbance vs. radius 2 /2. In this type of plot a single species gives a straight line whose slope is proportional to mass. The theoretical slope calculated for the monomer mass (~17 kda). The dark blue line (mostly hidden behind the data points) has the theoretical slope for the trimer mass. This plot therefore makes it obvious that this protein is indeed a trimer, and therefore it is indeed a homolog of TNF (and presumably is correctly folded).

Measuring binding affinities The function of many proteins is to bind to other proteins, and sedimentation equilibrium is a very powerful tool for studying such binding interactions. The graph at the right summarizes the data (points) and fitted curves for 8 experiments on mixtures of a monoclonal antibody and its ~25 kda protein antigen. The data sets cover experiments at different mixing ratios of antibody to antigen, and cover a wide range of concentrations. (Note that this entire set of experiments used only ~80 micrograms of antibody.) To analyze these data an appropriate binding model is needed. The model shown to the right is the simplest one possible for an antibody with two binding sites, and simply assumes that both sites have the same binding affinity and bind independently of one another (no cooperativity and no steric blocking of one site by antigen bound to the other). In fitting these data one is essentially asking: Is there a single value of the dissociation constant that can explain all 8 experiments? The solid lines in the graph above represent the best fit of this model, and the fact that the lines follow the data points quite well shows that this is a good fit.

Measuring sedmentation velocities In the centrifuge we have both a flux driven by the centrifugal force J = vc = ω 2 xsc, and a diffusive flux caused by concentration gradients that may be present: J = ω 2 xsc D( c/ x) Now the change in mass in a small slice of the rotor is dm/dt = JA(x) JA(x + dx), where the cross section A = axφ. The concentration change is the mass change divided by the volume between the surfaces: dc/dt = [1/(φxadx)]{JA(x) JA(x + dx)} = ( 1/φ xa)(dja/dx) = (1/x)(d(xJ)/dx Combining these two equations yields the Lamm equation for the ultracentrifuge: dc/dt = D [ (d 2 c/dx 2 ) + (1/x)(dc/dx) ] sω 2 [x(dc/dx) + 2c] (Note: we have ignored buoyancy, which opposes the centrifugal force; see pp. 354-355.)

Sample data one would fit to the Lamm equation The graph shows scans across the centrifuge cell, recording the absorbance at 280 nm versus position within the cell. These scans were taken starting at 13 minutes after initiating a run at 45,000 rpm (the black data set in the graph), and then every ~12 minutes thereafter (blue, green, cyan, etc.). In the first data set the sedimentation of the antibody has already depleted its concentration at the left and formed a sedimentation boundary. At later times in the run the depleted region expands and the boundary moves away from the center of the rotor, until by the time of the last data set the concentration of antibody has dropped to essentially zero throughout the upper half of the cell.

Some common data manipulations What we often want to know is how much material is sedimenting at various sedimentation coefficients. By taking many scans close together in time, subtracting them in pairs, and doing some mathematical manipulation these data can be transformed into the sedimentation coefficient distribution, g(s*), which is shown at the right. This distribution resembles a chromatogram, and the area under each peak gives the total amount of that species. For this antibody sample we see only one distinct peak, centered at a sedimentation coefficient of ~6.5 S, which corresponds to the native antibody monomer. A sedimentation coefficient of 6.5 S is actually rather low for a 150 kda species, which is consistent with high hydrodynamic friction from its highly asymmetric, non-globular Y shape. The red curve is a fit of these data as a single species. This fit clearly fails to account for the data over the region from 8-12 S, indicating the presence of some dimer and possibly also some trimer.

Still more data analysis

A microscopic picture of friction and diffusion We can write a microscopic description of a macromolecule encountering collisions with an (implicit) solvent as a Langevin equation: m d 2 ( ) x V dt 2 = f x ( ) dx + R(t) (6) dt The random force somehow has to be determined. It s clear that < R(t) >= 0, but beyond that, we will keep it undetermined right now. Let s see how quickly velocity fluctuations decay, starting with a free particle, V = 0, where the Langevin equation is m(dv/dt) = fv + R(t). Mutliplying by v(0) and averaging over many collisions: m v(0) dv(t) = f v(0)v(t) + v(0)r(t) dt The final term vanishes, so d dt v(0)v(t) = (f /m) v(0)v(t) v(0)v(t) = v 2 (0) e ft/m (7)

Kubo-Green relations The mean KE, 1/2m v 2 is given by 1/2kT. Now integrate over all times t: v(0)v(t) dt = kt m e ft/m dt = kt f = D 0 This is a Green-Kubo relation, connecting an equilibrium time-correlation function to a transport coefficient. A similar relationship is 0 0 R(0)R(t) dt = 2kTf = R 2 (0)