Rubber elasticity. Marc R. Roussel Department of Chemistry and Biochemistry University of Lethbridge. February 21, 2009

Similar documents
The lattice model of polymer solutions

VIII. Rubber Elasticity [B.Erman, J.E.Mark, Structure and properties of rubberlike networks]

Lattice protein models

Weight and contact forces: Young's modulus, Hooke's law and material properties

Chapter 3 - First Law of Thermodynamics

Chapter 3 Entropy elasticity (rubbery materials) Review basic thermal physics Chapter 5.1 to 5.5 (Nelson)

3 Biopolymers Uncorrelated chains - Freely jointed chain model

Exam Question 10: Differential Equations. June 19, Applied Mathematics: Lecture 6. Brendan Williamson. Introduction.

3.091 Introduction to Solid State Chemistry. Lecture Notes No. 6a BONDING AND SURFACES

PHYSICS 715 COURSE NOTES WEEK 1

Lecture 1 NONLINEAR ELASTICITY

Physics 3 Summer 1989 Lab 7 - Elasticity

Appendix 4. Appendix 4A Heat Capacity of Ideal Gases

First Law Limitations

Van der Waals Fluid Deriving the Fundamental Relation. How do we integrate the differential above to get the fundamental relation?

Chapter 20 Entropy and the 2nd Law of Thermodynamics

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:

Lecture 8. The Second Law of Thermodynamics; Energy Exchange

Chemistry 2000 Lecture 9: Entropy and the second law of thermodynamics

F = m a. t 2. stress = k(x) strain

Chapter 6: The Rouse Model. The Bead (friction factor) and Spring (Gaussian entropy) Molecular Model:

Lab Week 4 Module α 1. Polymer chains as entropy springs: Rubber stretching experiment and XRD study. Instructor: Gretchen DeVries

Physics is time symmetric Nature is not

Lecture 8. The Second Law of Thermodynamics; Energy Exchange

Homework Week 8 G = H T S. Given that G = H T S, using the first and second laws we can write,

2 Equations of Motion

Introduction Statistical Thermodynamics. Monday, January 6, 14

Chapter 2. Rubber Elasticity:

Boltzmann Distribution Law (adapted from Nash)

V(φ) CH 3 CH 2 CH 2 CH 3. High energy states. Low energy states. Views along the C2-C3 bond

1 Nonlinear deformation

2 Introduction to mechanics

Lecture Outline Chapter 17. Physics, 4 th Edition James S. Walker. Copyright 2010 Pearson Education, Inc.

Rubber Elasticity (Indented text follows Strobl, other follows Doi)

Assignment 3. Tyler Shendruk February 26, 2010

Lecture 7: Kinetic Theory of Gases, Part 2. ! = mn v x

Figure 1: Doing work on a block by pushing it across the floor.

REVIEW : INTERATOMIC BONDING : THE LENNARD-JONES POTENTIAL & RUBBER ELASTICITY I DERIVATION OF STRESS VERSUS STRAIN LAWS FOR RUBBER ELASTICITY

1 Implicit Differentiation

3.012 PS Issued: Fall 2003 Graded problems due:

5.60 Thermodynamics & Kinetics Spring 2008

Summary of last part of lecture 2

Section 20: Arrow Diagrams on the Integers

Mass on a Horizontal Spring

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Statistical Physics I Spring Term 2013 Notes on the Microcanonical Ensemble

THE SECOND LAW OF THERMODYNAMICS. Professor Benjamin G. Levine CEM 182H Lecture 5

Physics 172H Modern Mechanics

Lecture Notes Set 3b: Entropy and the 2 nd law

X α = E x α = E. Ω Y (E,x)

Rouse chains, unentangled. entangled. Low Molecular Weight (M < M e ) chains shown moving past one another.

Irreversible Processes

3.091 Introduction to Solid State Chemistry. Lecture Notes No. 5a ELASTIC BEHAVIOR OF SOLIDS

4.1 Constant (T, V, n) Experiments: The Helmholtz Free Energy

Soft Bodies. Good approximation for hard ones. approximation breaks when objects break, or deform. Generalization: soft (deformable) bodies

Chap. 2. Polymers Introduction. - Polymers: synthetic materials <--> natural materials

Chemistry 2000 Lecture 15: Electrochemistry

MA1131 Lecture 15 (2 & 3/12/2010) 77. dx dx v + udv dx. (uv) = v du dx dx + dx dx dx

Work and heat. Expansion Work. Heat Transactions. Chapter 2 of Atkins: The First Law: Concepts. Sections of Atkins

Building your toolbelt

Conservation of mass. Continuum Mechanics. Conservation of Momentum. Cauchy s Fundamental Postulate. # f body

FUNDAMENTALS OF THERMODYNAMICS AND HEAT TRANSFER

Properties of Entropy

Physics of Continuous media

Solutions to Problem Set 4

Proteins polymer molecules, folded in complex structures. Konstantin Popov Department of Biochemistry and Biophysics

Chapter 20 The Second Law of Thermodynamics

Physics 2B Chapter 17 Notes - First Law of Thermo Spring 2018

Physics Sep Example A Spin System

Lecture 14. Entropy relationship to heat

Chemical Engineering 160/260 Polymer Science and Engineering. Lecture 14: Amorphous State February 14, 2001

PHASE TRANSITIONS IN SOFT MATTER SYSTEMS

Chapter 5 - Systems under pressure 62

MECHANICAL PROPERTIES OF MATERIALS

Experimental Soft Matter (M. Durand, G. Foffi)

Thermodynamics (Classical) for Biological Systems Prof. G. K. Suraishkumar Department of Biotechnology Indian Institute of Technology Madras

Foundations of Chemical Kinetics. Lecture 12: Transition-state theory: The thermodynamic formalism

THERMODYNAMICS Lecture 2: Evaluating properties

Swelling and Collapse of Single Polymer Molecules and Gels.

Chapter 1. Basic Concepts. 1.1 Trajectories

Calculus 2 - Examination

Why Complexity is Different

vs. Chapter 4: Standard Flows Chapter 4: Standard Flows for Rheology shear elongation 2/1/2016 CM4650 Lectures 1-3: Intro, Mathematical Review

Chapter Two: Mechanical Properties of materials

Lecture 13. Multiplicity and statistical definition of entropy

Physics 141 Rotational Motion 2 Page 1. Rotational Motion 2

Phase Transitions. µ a (P c (T ), T ) µ b (P c (T ), T ), (3) µ a (P, T c (P )) µ b (P, T c (P )). (4)

Rubber Elasticity. Scope. Sources of Rubber. What is Rubber? MSE 383, Unit 2-6

Quiz 3 for Physics 176: Answers. Professor Greenside

Molecular Driving Forces

Class 13: Classical Particles and Quantum Particles

Entropy and the Second Law of Thermodynamics

Introduction. Statistical physics: microscopic foundation of thermodynamics degrees of freedom 2 3 state variables!

Current. Lecture 10. Chapter Physics II. Course website:

Lecture 9 Examples and Problems

Applied Thermodynamics for Marine Systems Prof. P. K. Das Department of Mechanical Engineering Indian Institute of Technology, Kharagpur

More Thermodynamics. Specific Specific Heats of a Gas Equipartition of Energy Reversible and Irreversible Processes

Magnets and Potential Energy

Introduction to Thermodynamic States Gases

Polymer Dynamics and Rheology

Module 2 : Electrostatics Lecture 9 : Electrostatic Potential

Transcription:

Rubber elasticity Marc R. Roussel Department of Chemistry and Biochemistry University of Lethbridge February 21, 2009 A rubber is a material that can undergo large deformations e.g. stretching to five or ten times its original length) and then return to its original shape and size. Many polymeric materials, and not just rubber itself, display rubber elasticity in some range of temperatures. When we characterize these materials, we find that they are made of long chains, with occasional cross-links between the chains. In the resting state, these chains will adopt random coil configurations. As we apply a stress by pulling on a sample of rubber, the chains will tend to align along the stress, which allows the material to stretch. The cross-links pull the material back to its original shape once the stress is removed. In this note, we will look at a statistical treatment of rubber elasticity. This will allow us to relate the elastic behavior to properties of the polymer. 1 Cross-linking statistics Figure 1 shows two primary polymer molecules which have formed some cross links. There are both internal cross-links within a single molecule) and external cross-links between molecules). Both types can be important to rubber elasticity. The important statistic for rubber elasticity is in fact the number of chains bounded by two cross-link junctions, a quantity called the number of active chains, denoted ν e. Free ends are uninteresting because they can be displaced more-or-less freely. If we have N primary polymers, there will be 2N free ends. Let ν be the number of monomers involved in cross-links two per cross-link). Each cross-link takes two chains and divides them into four, so the number of chains added by making ν/2 cross-links is equal to ν. From this number, we have to subtract the number of chains that end freely. Thus, ν e = ν 2N = ν 1 2N/ν). Now let M be the average molar mass of the primary polymers, and M c be the molar mass per cross-linked monomer, i.e. M c = NM/ν. Another way to look at M c is that it s the average molar mass of the active chains. We can then rewrite the last equation ν e = ν 1 2M c /M). 1) 1

Figure 1: Two primary polymer molecules i.e. two polymer chains) which have formed some cross-links, shown as bold dots. The chains are colored differently for ease of visualization. 2 Thermodynamics of elasticity When we write down the differential of U in courses in chemical thermodynamics, we normally only consider pressure-volume work. Here, we need to also consider the extension work. Suppose that f is the externally imposed equilibrium, i.e. reversible) tension on an elastic body. Then dw = P dv + f dl, where dl represents a change in the length of the body. You will recognize the new term as a special case of the general definition of mechanical work as force times distance. Then we have du = dw + dq = P dv + f dl + T ds. In this form, we have a differential of U in terms of the variables S, V, L). Physically, these are not the most convenient variables. We would like to rewrite U as a function of T, V, L). To do this, we need to rewrite the differential of S in terms of those variables: ds = S T dt + S V,L V dv + S T,L L dl. If we now substitute this equation into the differential of U, we get du = T S ) V P dv + T S ) T,L L + f dl + T S T dt. V,L U L = T S L + f, or f = U L T S L. 2) 2

If we can develop a theory for the thermodynamic properties of rubbers, then equation 2 will give us an equation for the force which must be applied to get a certain extension L from the equilibrium length L. The elasticity of rubbers is largely due to conformational rearrangements of the chains, which are obtained largely by rotations around σ bonds of the framework. These rotations should take very little energy. If we assume that they take no energy at all, then we get the ideal rubber model. In terms of the quantities in equation 2, an ideal rubber is one for which U L = 0. The elastic force is then due to an entropic effect alone: f = T S L. 3) It is worthwhile spending a few minutes thinking about what equation 3 tells us in relation to rubber elasticity. When we pull on a piece of rubber, we straighten out the polymer chains. This decreases the entropy, so S/ L < 0. Another way to put this is that there are more microstates associated with the collapsed state than with the stretched state. The random thermal motions of the chains in a stretched state are therefore more likely to be in a direction which would tend to bring the polymer back toward its relaxed state, which generates the resistive force f. This is remarkable: Entropy, which we normally think of as a somewhat abstract statistical quantity, is directly responsible for a macroscopic force! 3 Stress and strain In materials science, we don t normally talk about the force vs extension relationship, which is what equation 3 would give us. There are two problems: 1. The force depends on the geometry of the sample, in particular its cross-sectional area perpendicular to the applied force. To solve this problem, we define the stress σ = F/A. 2. The actual extension L) for a given stress depends on the initial length of the sample, L 0. On the other hand, the strain ε = L/L 0 = L L 0 )/L 0 only depends on the stress. Equivalently, we can use the stretch ratio λ = L/L 0 = ε + 1. We will use λ in these notes, although clearly λ and ε are interchangeable quantities. 3

Materials scientists thus usually discuss deformation using the stress-strain relationship. The term stress-strain relationship is used even when the strain is described in terms of the stretch ratio.) From the definition of λ, we have L = λl 0. Therefore L = 1 L 0 λ. Equation 3 therefore becomes f = T L 0 S λ. 4) 4 Statistical theory of rubber elasticity To calculate the entropy of a rubber, we will use the Boltzmann equation: S = k ln W, where W is the number of microstates. The active chains may have different lengths, which affects W. We can compute W by W = n W n ) νn, where ν n is the number of chains containing n bonds. This gives S = k n ν n ln W n. We need to reinterpret this equation a little: In our case, we can t count the microstates. Suppose that I give you the probability density p n λ) that a given n-mer has end-to-end is stretched by a factor λ from its resting length. The number of microstates between λ and λ+dλ is proportional to p n λ) dλ. Since we are going to take a logarithm, the proportionality constant, as well as the factor of dλ, end up in an additive constants: S = k n ν n ln p n λ) + constant). Our next step will be to take the derivative in equation 4, so the constant will vanish: f = kt ln p n λ) ν n L 0 λ. 5) n Cross-links occur where two primary polymers have come sufficiently close together in the right orientation. Cross-linking should therefore have only a small effect on the locations of the monomers which became cross-linked. The distribution of relative positions of the two 4

cross-link junctions defining an active chain should therefore obey the Gaussian probability density derived in the last lecture: p n x, y, z) = q/ π) 3 e q2 x 2 +y 2 +z 2), 6) where q 2 = 3/2nl 2 ). We also had r 2 0 = nl 2 = 3/2q 2 ), where r 2 0 is the mean squared end-to-end distance of the relaxed polymer. We now introduce an affine transformation: We assume that the deformation of a sample of rubber stretches all the coordinates homogeneously by factors λ x, λ y and λ z in each of the three Cartesian directions. Thus, x, y, z) = λ x x 0, λ y y 0, λ z z 0 ), where x 0, y 0, z 0 ) are the initial relative coordinates of the end of a chain. Taking a logarithm of equation 6 and introducing the affine transformation, we get ln p n = q 2 [ λ x x 0 ) 2 + λ y y 0 ) 2 + λ z z 0 ) 2] + constant, where again we won t be too concerned with the constants since they won t contribute to the force. Rubbers are essentially incompressible, so we must have λ x λ y λ z = 1. Suppose that we stretch a sample of rubber along the x axis. Let λ x = λ. We will typically find that λ y = λ z. 1 Substituting this relationship into the incompressibility condition, we get λ y = λ z = 1/ λ. This gives ln p n λ) = q 2 [ λ 2 x 2 0 + y 2 0 + z 2 0)/λ ] + constant. There is just one more thing: If we have ν n polymers of length n, then we should average this quantity over all the initial end-to-end distances: ln p n λ) = q 2 [ λ 2 x 2 0 + y 2 0 + z 2 0 )/λ ] + constant. For the relaxed polymer, there is nothing special about the x axis, i.e. x 2 0 = y 2 0 = z 2 0. Since r 2 0 = x 2 0 + y 2 0 + z 2 0, this gives x 2 0 = y 2 0 = z 2 0 = r 2 0 /3. We therefore obtain ln p n λ) = q2 r 2 0 3 λ 2 + 2/λ ) + constant = 1 2 λ 2 + 2/λ ) + constant. Note that q and r0 2 have both dropped out of the non-constant part of this expression. The derivative in equation 5 therefore doesn t depend on n: ln p n λ) L = λ 1/λ 2). It can therefore be pulled out of the sum, leaving n ν n = ν e, the number of active chains. Putting it all together, we get the elastic force f = kt ν e L 0 λ 1/λ 2 ). 1 Some polymers do deform differently in different directions, but then we would have to orient the sample correctly with respect to any special axes to observe this phenomenon. 5

To get the strain, we divide this equation by the cross-sectional area. There are actually two slightly different conventions in use here. One is to divide by the cross-sectional area of the relaxed sample, A 0, giving σ 0 = kt ν e λ 1/λ 2 ). 7) σ 0 is called the nominal stress. This is the convention normally used by polymer scientists. The other is to divide by the area corresponding to stretch ratio λ. Since the y and z are transformed by the ratio λ y = λ z = 1/ λ, A = A 0 /λ. The true stress is therefore σ = kt ν e λ 2 1/λ ). We will pursue our analysis with equation 7. Using equation 1, the stress becomes σ 0 = kt ν 1 2M c /M) λ 1/λ 2). If we define the number of moles of cross-linked monomers n c = ν/n A, we get σ 0 = n crt 1 2M c /M) λ 1/λ 2). Almost the entire volume is occupied by active chains. The mass of the sample is therefore approximately n c M c. If we denote by v the specific volume of the relaxed sample, then = n c M c v, and therefore σ 0 = RT v 1 2 ) λ ) 1/λ 2. M c M This is our final expression. It predicts two interesting properties of the stress-strain relationship: 1. For a given strain, the stress increases linearly with temperature. Alternatively, if we fix the stress e.g. by suspending a mass from a piece of rubber), then the extension should decrease as we increase the temperature. This is a well known property of rubber which the statistical theory correctly predicts. Note that this behavior is the opposite of what we observe for normal materials: Typically, heating something up causes it to expand. Putting a stress on normal materials e.g. metals) as they are being heated does not change this behavior. 2. Increasing the number of cross-links, keeping all other properties constant, decreases M c, which in turn increases the stress at fixed strain. This is probably intuitively obvious, but it s nice that the statistical theory actually predicts this behavior correctly. 6