Вы находитесь на странице: 1из 26

MM 538 Engineering Rheology/Viscoelasticity Rheology Science of deformation and flow of matter

Dr. Kausar Ali Syed Polymer Lecture No 8, 9, 10, & 11 Oct 6, 13, & 27 2010

1. Hookean elasticity: Motion of chain segments is drastically restricted. Only bond stretching and bond angle deformation. Glassy behaviour 2. Viscous flow: 3. Viscoelasticity: The irreversible bulk deformation of polymeric material (irreversible slippage of molecular chains) Deformation is reversible but time dependent and associated with the distortaion of polymer chains through activated segment motion involving rotation about chemical bonds. Local freedom of motion associated with small-scale movement of chain segment is retained, but large movement (flow) is prevented by the restraint of a diffuse network structure.

4. Rubber elasticity:

In polymers the bonds between adjacent macromolecules are comparatively weak secondary bonds. In addition the tangled macromolecules can accommodate deformation by uncoiling. Thus polymers have much lower elastic moduli than crystalline metals and ceramics. Thermoplastic material below their glass transition temperature Tg and thermoset polymers resist sliding motion between macromolecules. These deformation characteristics are not significantly influenced by temperature; therefore, the elastic moduli of thermoplastics below Tg and crosslinked polymers do not change significantly with temperature. For relatively small deformations, the mechanical behavior at low temperatures may be elastic. Elastic deformation is instantaneous, which means that total deformation (or strain) occurs the instant the stress is applied or released. In other words we say that the response of amorphous material below Tg to an applied force is independent of time. Also, such deformation is fully recoverable upon removal of stress, and it follows a stress-strain relationship similar to the one followed by metals and ceramics, i.e., the strain is always proportional to the stress and is independent of the rate at which the body is deformed: = E (1)

where E is Youngs modulus if the deformation mode is a tensile stretch and equation (1) is an expression of famous Hookeans law. Conversely, the strain and stress are related by the tensile compliance, D, defined by = D (2)

For tensile deformation, the compliance is therefore the reciprocal of the modulus D = 1/E (3)

Stretching of primary covalent bonds and some early stages of uncoiling of long chain molecules: Not much temperature dependent & fully recoverable upon removal of stress. Equation (1) summarized purely elastic response in tension. The analogous expression for shear deformation can be obtained as explained below. Consider a rectangular block of material with a top surface area A, and height dx, subject to a shearing force F, as shown in the fig. 1 (a). When the block is a solid, the shearing force is directly proportional to the deformation dy (i.e., F dy). To make this relationship independent of sample dimension , the force F is divided by the area A, and the displacement dy is divided by the height dx, yielding the relationship:

(F/A) dy/dx where F/A is called the shear stress, , and dy/dx as the shear strain, , and the equation is written as: or = G (4)

where G is a constant known as shear modulus.

Fig. 1 At the highest temperatures, where there is enough molecular motion, viscous or liquid-like behavior prevails or the deformation is viscous deformation. When the solid in the above example is replaced with a fluid (fig. 1 b), the shearing force no longer produces a characteristic displacement (strain). Instead, the displacement continues with time, i.e., the response of amorphous material above Tg is time-dependent. The shearing force is proportional to the rate of displacement, d(dy/dt), i.e., F d(dy/dt). Normalizing the force by the area and the displacement rate by the height yields the expression: d(dy/dt)/dx.

Since dx is a constant, we can rearrange the equation to give: d(dy/dx)/dt Recognizing that dy/dx is shear strain , the equation becomes: or = d /dt d/dt (5)

where the constant is known as the viscosity, which is a measure of work done on a material undergoing shear deformation. This equation is known as Newtons law of viscosity. Viscosity is an important property and the study of deformation and flow of materials is called rheology. The units of viscosity are poise, P (= g/cm/sec ). When the temperature in the thermoplastic materials exceeds Tg, molecular motion can occur and can cause viscous deformation. The extent of viscous deformation increases rapidly with temperature, causing the elastic modulus to decrease rapidly with increasing temperature. Since above Tg, viscous deformation occur simultaneously with elastic deformation, the behavior in this regime is known as Viscoelasticity. Viscoelasticity is the property of materials that exhibit both viscous and elastic characteristics when undergoing deformation. Viscous materials, like honey, resist shear flow and strain linearly with time when a stress is applied. Elastic materials strain instantaneously when stretched and just as quickly return to their original state once the stress is removed. Viscoelastic materials have elements of both of these properties and, as such, exhibit time dependent strain. Whereas elasticity is usually the result of bond stretching along crystallographic planes in an ordered solid, viscosity is the result of the diffusion of atoms or molecules inside an amorphous material.

Specifically, viscoelasticity is a molecular rearrangement. When a stress is applied to a viscoelastic material such as a polymer, parts of the long polymer chain change position. This movement or rearrangement is called Creep. Polymers remain a solid material even when these parts of their chains are rearranging in order to accompany the stress, and as this occurs, it creates a back stress in the material. When the back stress is the same magnitude as the applied stress, the material no longer creeps. When the original stress is taken away, the accumulated back stresses will cause the polymer to return to its original form. The material creeps, which gives the prefix visco-, and the material fully recovers, which gives the suffix -elasticity. Both Hookes and Newtons laws are valid for small changes in strain and rate of strain, and both are useful in studying the effect of stress on viscoelastic materials. The initial elongation of a stressed polymer below Tg is the reversible elongation due to a stretching of covalent bonds and distortion of the bond angles. Some of the very early stages of elongation by disentaglement of chains may also be reversible. However the rate of flow, which is related to slower disentaglement and slippage of polymer chains past each other, is irreversible and increases ( decreases) as the temperature increases in accordance with the following Arrehenius equation in which E is the activation energy for viscous flow. = A e E/RT (6)

The viscous portion is also time dependent. In other words if the loading rate is high, less time is available for time dependent deformation and the viscous behavior is suppressed and the elastic moduli is higher than for lower loading rate. The cause of the stress relaxation is that viscous flow in the polymeric materials internal structure occurs by the polymer chains slowly sliding past each other by the breaking and reforming of secondary bonds between the chains and by mechanical untangling and recoiling of the chains. Stress relaxation allows the material to attain a lower energy state spontaneously, if there is sufficient activation energy for the process to occur. Stress relaxation of polymeric materials is thus temperature dependent and associated with an activation energy.

Fig. 2 The stress relaxation of a stressed polymeric material under constant strain (fig. a) results in a decrease in the stress with time (fig. b).

The response of different materials-ideal elastic, ideal viscous, and viscoelastic to a step change in stress (or load) is illustrated in figure 3. For an ideal (Hookean) elastic materials, the resulting strain is instantaneous and constant during the duration of the applied stress. When the load is removed at t = t r, the strain instantaneously drops to zero (figure 3b). This follows from Hookes law applied to the case of a constant stress: = D 0 = 0 (7)

In the case of an ideal (Newtonian) viscous fluid, the strain response is obtained by integration of Newton law of viscosity ( = d /dt) arranged as

= ( )

= (0 / ) t

(8)

Fig. 3 Here the strain increase with increasing time until the load is removed at t = t r at which time the strain remains constant (0 tr / ) in time (fig. 3d) unless an additional load is applied. The deformation is said to be permanent since no elastic recovery is possible. In the case of a viscoelastic material, recovery can occur due to the elastic contribution of the materials, the strain response will have some of the character of both elastic and viscous materials as shown in figure 3c. The initial portion of the strain recovery is elastic (i.e. instantaneous) while full recovery is delayed to longer times due to the viscous contribution. The actual viscoelastic can be responsibly modeled by analyzing the strain response of series or parallel combination of ideal elastic springs and ideal viscous dashpots (shock absorber) as discussed later. Phenomenological Viscoelasticity Three types of experiments are used in the study of viscoelasticity which can be used to characterize the dimensional stability of a material. These involve: Time Dependent Increase in Strain under Constant Stress known as Creep (Time Dependent Deformation) Time Dependent Reduction in Stress at Constant Strain known as Stress Relaxation. Dynamic techniques (periodical variation of stress or deformation instead of a constant load or strain)

Fig. 4. Deformation of various polymer types when stress is applied and unloaded, (a) Crosslinked ideal elastomer, (b) Fiber, (c) Amorphous plastic.

In creep studies a body is subjected to a constant stress 0 and the sample dimensions are monitored as a function of time. When the polymer is first loaded an immediate deformation occurs, followed by progressively slower dimensional changes as the sample creeps towards a limiting shape. Figure 3 shows examples of the different behaviors observed in such experiments. Stress relaxation is an alternative procedure. Here an instantaneous, fixed deformation 0 is imposed on a sample at constant temperature, and the stress decay is followed with time. A very useful modification of these two basic techniques involves the use of a periodically varying stress or deformation instead of a constant load or strain. The dynamic responses of the body are measured under such conditions Creep test can be made in shear, torsion, flexure, or compression modes, as well as in tension. Results of these tests are particularly important for selecting a polymer that must sustain loads for long period. Consider the tensile experiment as a creep study in which a steady stress 0 is suddenly applied to the polymer specimen. In general, the resulting strain (t) will be a function of time starting from the imposition of the load. Usually, the parameter of interest is the tensile compliance, D, which is the ratio of strain to stress. While the stress, 0, is held constant during the creep test, the resulting strain (t) depends upon the time during which the load has been applied and, therefore, the tensile creep compliance D(t) also becomes a function of time as D (t) = (t) / 0 The shear creep compliance J(t) (see Fig. 1b in the box) is similarly defined as J(t) = (t) /0 where 0 is the constant shear stress and (t) is the resulting time-dependent strain. In stress relaxation experiments, a rapid extension (0) sketched in Fig. 1 (box) is applied to the sample and held constant while the resulting stresses on the sample is measured as a function of time and, therefore, the tensile stress relaxation modulus, Er, is also time dependent. Er(t) = (t) / 0 (11) (10) (9)

with 0 being the constant strain. Similarly, a shear relaxation experiment measures the shear relaxation modulus G(t): G(t) = (t)/0 (12)

where 0 is the constant strain. Although a compliance is the inverse of a modulus for an ideal elastic body (i.e., D = 1 / E) this is not true for viscoelastic materials. That is, Er(t) = (t)/o 1 /[(t)/0] = 1 / D(t) and G(t) = (t)/ 0 1 /[(t) /0] = 1 / J(t) (14) (13)

Consider two experiments carried out with identical samples of a viscoelastic material. In experiment (a) the sample is subjected to a stress 1 for a time t. The resulting strain at t is 1, and the creep compliance measured at that time is D1(t) = 1/1. In experiment (b) a sample is stressed to a level 2 such that strain 1 is achieved immediately. The stress is then gradually decreased so that the strain remains at 1 for time t (i.e., the sample does not creep further). The stress on the material at time t will be 3, and the corresponding relaxation modulus will be E2 (t)=3/1. In measurements of this type, it can be expected that 2 > 1 > 3, and E (t) (D(t))-1 as indicated in Eq. (11). E(t) is obtained directly only from stress relaxation measurements, while D(t) require creep experiments for its direct observation. These various parameters can be related in the linear viscoelastic region. Mechanical Models of Viscoelastic Behavior An insight into the nature of polymers can be obtained by analyzing the stress or strain response of mechanical models using an ideal spring with a modulus of E as the Hookean element and a dashpot as the viscous element. A dashpot may be viewed as a shock absorber consisting of a piston in a cylinder filled with a Newtonian fluid of viscosity . These models demonstrate the deformations of an elastic solid and an ideal liquid. The stress-strain curves for these models are shown in fig. 2. In general terms, the Hookean spring represents bond flexing while the Newtonian dashpot represents chain and local segmental movement. Again, in general terms, below their Tg polymers can be modeled as having a behavior where the spring portion is more important. Above their Tg where segmental mobility occurs, the dashpot portion is more important. Since polymers are viscoelastic solids, combination of these models are used to demonstrate the deformation resulting from the application of stress to an isotropic solid polymer. Mechanical models are useful tools for selecting appropriate mathematical functions to describe particular phenomena. The models have no physical relation to real materials, and it should be realized that an infinite number of different models can be used to represent a given phenomenon. Two models are mentioned here to introduce such concepts, which are widely used in studies of viscoelastic behavior. The elemental models are a series combination of a spring and a dashpot, the Maxwell element, and a parallel combination of a spring and a dashpot, a Voigt element. The Maxwell body is appropriate for the description of stress relaxation, while the Voigt element is more suitable for creep deformation. Maxwell element This model consists of a spring and dashpot in series, as shown schematically in fig. 5, the total strain (or strain rate) is a summation of the individual strains (or strain rates) of the spring and dashpot. If a fixed strain is suddenly applied, the spring responds immediately by extending and a stress is produced in it, which is therefore also applied to the dashpot. The dashpot cannot be displaced instantaneously but begins to be displaced at a rate proportional to the stress. Thus, the application of stress would cause an instantaneous elongation of the spring, followed by a slow response of the piston in the dashpot. The strain and stress in the spring thus decay to zero as the dashpot is displaced at a decreasing rate and is eventually displaced by the same amount as the spring was originally displaced. This is therefore a model for stress-relaxation, but the stress relaxes to zero, which is not always the case for real polymers. Under

constant stress, the spring remains at constant length, but the dashpot is displaced at a constant rate, so that the model cannot describe creep.

Fig. 5

The Maxwell model: spring and dashpot in series.

Stress - Strain Relations The spring is the elastic component of the response and obeys the relation s = Es where (s and s are the stress and strain respectively and E is a constant. The dashpot is the viscous component of the response and in this case the stress d is proportional to the rate of strain dd /dt, ie d = dd /dt where is a material constant. Equilibrium Equation For equilibrium of forces, assuming constant area Applied Stress, = s = d Geometry of Deformation Equation The total strain, is equal to the sum of the strains in the two elements. so = s + d Thus from equations 27, 28, and 30 d / dt or = ds / dt + dd /dt (19) (17) (16) (15)

(18)

This is the governing equation of the Maxwell Model. It is interesting to consider the response that this model predicts under three common time dependent modes of deformation. This differential equation can be solved for creep and stress relaxation response by applying the appropriate stress or strain function. (i) Creep In a creep experiment, a constant stress, 0, is applied instantaneously. Eq.28 then reduces to

d / dt

= 0 /

(20)

which indicates a constant rate of increase of strain with time. Rearrangement and integration of eq. 32 gives (t) = [0 / ] t + 0 (21)

where 0 (= 0/E) represents the instantaneous (i.e., t = 0) strain response of the spring element. From Fig. 6 it may be seen that for the Maxwell model, the strain at any time, t, after the application of a constant stress, 0, is given by eq. 33, i.e., (t) = [0 / ] t + 0

Fig. 6 Response of Maxwell model

The creep compliance function , D(t), is then obtained as D(t) = (t) / 0 = t / + 0 / 0 = t / + D where D (=0/0) is the instantaneous compliance (of the spring). Using the relation = /E, an alternative form of eq. 22 is D(t)/ D = t / + 1 And, the creep modulus, Ec(t), is given by E(t) (ii) Relaxation For stress-relaxation, strain is held constant, ( is independent of t), so that strain rate d /dt = 0 and 1/E (d/dt) + / = 0 (24) = 0 / (t) = (23) (22)

Rearrangement of eq. 24 and introduction of gives d/ = (1 / ) dt where = /E is called a relaxation time Integration of eq. 37 with the initial condition = 0 at t = t0 , leads to (t) = 0 exp (E t / ) or with 0 = E, and (t) = 0 exp ( t / ) (26) (25)

(t) 0

= =

stress after time t initial stress (instantaneous stress response of the spring)

The rate at which stress relaxation occurs depends on the relaxation time , which is a property of the material and which is defined as the time needed for the strees () to decrease to 0.37 (1/e) of the initial stress 0. The stress decays exponentially to zero with a relaxation time = / E. It follows from equation (31) that, if is constant, d/dt = / and is constant. This is the formal expression of the fact that this model cannot describe creep, as has already been indicated. The stress relaxation modulus, Er(t), is Er(t) = / 0 = (0 / 0) exp ( t / ) = E exp ( t / ) where E (= 0/0) is the (Youngs) modulus of the spring element. (iii) Recovery When the stress is removed there is an instantaneous recovery of the elastic strain, ', and then, as shown by equation (28), the strain rate is zero so that there is no further recovery (see Fig. 6). It can be seen therefore that although the relaxation behaviour of this model is acceptable as a first approximation to the actual materials response, it is inadequate in its prediction for creep and recovery behaviour. (27)

The Kelvin or Voigt model This model consists of a spring and dashpot in parallel, as shown schematically in fig. 7. A simple creep experiment involves application of a stress 0 at time t = 0 and measurement of the strain while the stress is held constant. If a fixed stress is suddenly applied, the dashpot cannot be displaced instantaneously, so that the spring does not change in length and carries none of the stress. The dashpot is then displaced at a decreasing rate as the spring strains and takes up some of the stress. Eventually the dashpot and spring have both been displaced far enough for the spring to take the whole load. This is thus a model for creep. For constant strain, there is no relaxation of stress and there is also no way to apply an immediate finite strain, because this would require an infinite stress to be applied to the dashpot. The model cannot therefore describe stress-relaxation.

Fig. 7 Stress-Strain Relations These are the same as the Maxwell Model and are given by equations (27) and (28). Equilibrium Equation For equilibrium of forces it can be seen that the applied load is supported jointly by the spring and the dashpot, so = s + d (28)

Geometry of Deformation Equation In this case the total strain is equal to the strain in each of the elements, i.e. = s = d From equations (27), (28) and (40) = Es + dd /dt or using eq. 41 = E + d / dt (30) (29)

This is the governing equation for the Kelvin (or Voigt) Model and it is interesting to consider its predictions for the common time dependent deformations. (i) Creep If a constant stress, 0 is applied then equation (39) becomes 0 / = E / + d / dt or 0 / = (t) / + d(t) / dt (31)

where = /E is called a retardation time in a creep experiment. Equation (43) can be solved by using an integrating factor (et/). Integration from = 0, = 0 at t = 0 gives E / 0 = 1 exp (E t/) (t) = 0 / E [1 exp ( t / )] (32)

This indicates an exponential increase in strain from zero up to the value, 0 / E , that the spring would have reached if the dashpot had not been present. This is shown in Fig. 8.

Fig. 8 Response of Kelvin/Voigt model and the compliance function is given by D(t) = D [1 exp ( t / )] where D(t) = (t)/0 and D = 1/E The strain rises exponentially to 0/E, with retardation time = /E . (33)

And, as for the Maxwell model, the creep modulus, E(t) may be determined as E(t) (ii) Relaxation If the strain is held constant then equation (39) becomes = E. That is, the stress is constant and supported by the spring element so that the predicted response is that of an elastic material, i.e. no relaxation (see Fig. 11). The fact that this model cannot describe stress-relaxation is formally expressed by equation (39), which shows that, if d/dt = 0, = E and is constant. (iii) Recovery If the stress is removed, then equation (39) becomes 0 = E + d / dt (35) = 0 / (t) = E [1 et / ] 1 (34)

Solving this differential equation with the initial condition = at the time of stress removal, then (t) = eEt/ (36)

This represents an exponential recovery of strain which is a reversal of the predicted creep. The Kelvin element does not display stress relaxation under constant strain conditions and the Maxwell model does not exhibit full recovery of strain when the stress is removed. A combination of the two mechanical models can be used, however, to represent both the creep and stress relaxation behaviors of polymers. This is the standard linear solid, or Zener model, comprising either a spring in series with a Kelvin element or a spring in parallel with a Maxwell model. Limitations to the effectiveness of mechanical models occur because actual polymers are characterized by many relaxation times instead of single values and because use of the models mentioned assumes linear viscoelastic behavior which is observed only at small levels of stress and strain. The linear elements are nevertheless useful in constructing appropriate mathematical expressions for viscoelastic behavior and for understanding such phenomena. Linear Viscoelasticity Linear viscoelastic behavior is actually observed with polymers only in very restricted circumstances involving homogeneous, isotropic, amorphous specimens subjected to small strains at temperatures near or above Tg and under test conditions that are far removed from those in which the sample may be broken. Linear viscoelasticity theory is of limited use in predicting service behavior of polymeric articles, because such applications often involve large strains, anisotropic objects, fracture phenomena, and other effects which result in nonlinear behavior. The theory is nevertheless valuable as a reference frame for a wide range of applications, just as the thermodynamic equations for ideal solutions help organize the observed behavior of real solutions. The major features of linear viscoelastic behavior that will be reviewed here are the superposition principle and time-temperature equivalence. Where they are valid, both make it possible to calculate the mechanical response of a material under a wide range of conditions from a limited store of experimental information.

(i) Boltzmann Superposition Principle Intermittent Loading The creep behaviour of plastics considered to date has assumed that the level of the applied stress is constant. However, in service the material may be subjected to a complex pattern of loading and unloading cycles. This can cause design problems in that clearly it would not be feasible to obtain experimental data to cover all possible loading situations and yet to design on the basis of constant loading at the maximum stress would not make efficient use of material or be economical. In these cases it is useful to have methods of predicting the extent of the recovered strain which occurs during the rest periods of conversely the accumulated strain after N cycles of load changes. There are several approaches that can be used to tackle this problem and two of these will be considered now. Superposition Principle The simplest theoretical model proposed to predict the strain response to a complex stress history is the Boltzmann Superposition Principle. Basically this principle proposes that for a linear viscoelastic material, the strain response to a complex loading history is simply the algebraic sum of the strains due to each step in load. Implied in this principle is the idea that the behaviour of a plastic is a function of its entire loading history. There are two situations to consider. (a) Step Changes of Stress When a linear viscoelastic material is subjected to a constant stress, 1, at time, t1, then the creep strain, (t), at any subsequent time, t, may be expressed as (t) = [1 / Ec(t t1)].1 (37)

where E(t t1) is the time-dependent modulus for the elapsed time (t t1). Then suppose that instead of this stress 1, another stress, 2 is applied at some arbitrary time, t2, then at any subsequent time, t, the stress will have been applied for a time (t - t2) so that the strain will be given by (t) = [1 / Ec(t t2)].2 (38)

Fig. 9 A two-step creep experiment. See the text for explanation and symbols.

As an example, fig. 9 illustrates a two-step loading programme. Now consider the situation in which the stress, 1, was applied at time, t1, and an additional stress, 2, applied at time, t2, then Boltzmanns Superposition Principle states that the total strain at time, t, is the algebraic sum of the two independent responses. (t) = [1 / Ec(t t1)].1 + [1 / Ec(t t2)].2 (39)

This equation can then be generalised, for any series of N step changes of stress, to the form

(40) where i is the step change of stress which occurs at time, ti . To illustrate the use of this expression, consider the following example (See box 1). Box 1

Example: A plastic which can have its creep behaviour described by a Maxwell model is to be subjected to the stress history shown in Fig. below. If the spring and dashpot constants for this model are 20 GN/m2 and 1000 GNs/m2 respectively then predict the strains in the material after 150 seconds, 250 seconds, 350 seconds and 450 seconds.

Solution : For the Maxwell model, the strain up to 100s is given by (t) = / E + t / Also the time dependent modulus Ec(t) is given by Ec(t) = / (t) = E / ( + Et) (2.56)

Then using equation (2.54) the strains may be calculated as follows: (i) at t = 150 seconds; 1 = 10 MN/m2 at tl = 0, 2 = 10 MN/m2 at t2 = 100 s (150) = 1 *

+ + 2 *

= 0.002 0.001 = 0.1% (ii) at 250 seconds; 1, 2 as above, 3 = 5 MN/m2 at t3 = 200 s (250) = 1 *

+ + 2 *

10 *

+ + ( 10) *

+ + 5 *

= (iii)

0.003 - 0.002 + 0.0005 = 0.15%

at 350 seconds; 1, 2 , 3 as above, 4 = 10 MN/m2 at t4 = 300 s (350) = 0.003 = 0.3%

(iv) and in the same way (450) = 0.004 = 0.4% The predicted strain variation is shown in Fig. 2.43(b). The constant strain rates predicted in this diagram are a result of the Maxwell model used in this example to illustrate the use of the superposition principle. Of course superposition is not restricted to this simple model. It can be applied to any type of model or directly to the creep curves. The method also lends itself to a graphical solution as follows. If a stress 01 is applied at zero time, then the creep curve will be the time dependent strain response predicted by equation (2.54). When a second stress, 2 is added then the new creep curve will be obtained by adding the creep due to 2 to the anticipated creep if stress a1 had remained alone. This is illustrated in Fig. 14(a). Then if all the stress is removed this is equivalent to removing the creep strain due to 1 and 2 independently as shown in Fig. 14(b). The procedure is repeated in a similar way for any other stress changes.

(b) Continuous Changes of Stress If the change in stress is continuous rather than a step function then equation (40) may be generalised further to take into account a continuous loading cycle. So

(41) where (t) is the expression for the stress variation that begins at time, t1. The lower limit is taken as minus infinity since it is a consequence of the Superposition Principle that the entire stress history of the material contributes to the subsequent response. It is worth noting that in exactly the same way, a material subjected to a continuous variation of strain may have its stress at any time predicted by

(42) To illustrate the use of equation (42) consider the following example (See box 2).

Box 2 A plastic is subjected to the stress history shown in Fig. below. The behaviour of the material may be assumed to be described by the Maxwell model in which the elastic component E = 20 GN/m2 and the viscous component = l000 GNs/m2. Determine the strain in the material (a) after u1 seconds (b) after u2 seconds and (c) after u3 seconds.

Fig. Stress history to be analysed The modulus for a Maxwell element may be expressed as Ec (t) = or 1/Ec (t) = (See Foot Note)

(a) The stress history can be defined as < t < 0, (t) = 0 d (t) / dt = 0 0 < t < T, (t) = K1 t d (t) / dt = K1 Equation for the continuous change of stress is

Substituting in above equation

It is interesting to note that if K1 was large (say K1 = 10 in which case T = 1 second) then the strain predicted after application of the total stress (10 MN/m2) would be (1) 0.0505%. This agrees with the result in the previous Example in which the application of stress was regarded as a step function. The reader may wish to check that if at time T = 1 second, the stress was held constant at 10 MN/m2 then after 100 seconds the predicted strain using the integral expression would be (100) 0.1495% which again agrees with the previous example.

Example 2 In the previous example, what would be the strain after 125 seconds if (a) the stress remained constant at 10 MN/m2 after 100 seconds and (b) the stress was reduced to zero after 100 seconds.

Solution:

(b) If the stress was completely removed after 100 seconds as shown in Fig. below then the effective change in stress would be given by change in stress, (t) = K1(t - T) d (t) / dt = K1

Foot Note

= s = d and = s + d d / dt = ds / dt + dd /dt d / dt = 1/E (d/dt) + /

(1)

In a creep experiment, a constant stress, 0, is applied instantaneously. Equation 1 then reduces to d / dt = 0 / (2) Rearrangement and integration of equation 2 gives (t) = [0 / ] t + 0 (3) where 0 = the instantaneous (t = 0) strain response of the spring element. The creep compliance function , D(t) = 1 /Ec(t), is D(t) = 1/Ec(t) = (t) / 0 = t / + 0 / 0 = t / + 1/E = where (E = 0 /0) is the instantaneous elastic modulus (of the spring) 1/Ec (t) =

It is apparent therefore that the Superposition Principle is a convenient method of analysing complex stress systems. However, it should not be forgotten that the principle is based on the assumption of linear viscoelasticity which is quite inapplicable at the higher stress levels and the accuracy of the predictions will reflect the accuracy with which the equation for modulus (equation 23) fits the experimental creep data for the material. In Examples (Box 1 and Box 2) a simple equation for modulus was selected in order to illustrate the method of solution. More accurate predictions could have been made if the modulus equation for the combined Maxwell / Kelvin model or the Standard Linear Solid had been used. An interesting and important application of the BSP is creep followed by recovery. If the loaded specimen is allowed to elongate for some time and the stress is then removed, creep recovery will be observed. Consider the loading programme shown in fig. 4: apply stress at time t 1; hold the stress constant until time t2 and then reduce the stress to zero. This last step is equivalent to applying an additional stress at t2. Thus, according to the BSP, at t > t2, (t) = D(t t1) D(t t2). (43)

Note that recovery r is not defined as one might expect; it is defined as the difference between the existing strain at t and the strain that would have been observed at t if the stress had not been removed.

Fig. 10 Creep and recovery. The upper graph shows the applied stress as a function

Creep and stress-relaxation are different manifestations of the viscoelastic nature of a polymer and must be related to each other. It is possible to show that = t (44)

Time Temperature Superposition Often, it is important to know how a material will behave (e.g., creep or stress relaxation) at a fixed temperature, but over a long time period (perhaps years) that may not be realistically accessible. Fortunately, long time behavior can be evaluated by measuring stress-relaxation or creep data over a shorter period of time but at several different temperatures. Information from each of these different temperature curves may then be combined to yield a master curve at a single temperature by horizontally shifting each curve along the log time scale. This technique is called time-temperature superposition and is a foundation of linear viscoelasticity theory. In this procedure, the master curve is plotted as stressrelaxation modulus or creep compliance versus reduced time, t/aT. The shift factor, aT, is defined as the ratio of (real) time to reach a particular value of modulus at some temperature to the reference-scale time coordinate, tr, corres[ponding to the same value of modulus in the master curve at the reference temperature, Tr, a T = t / tr

(45) Stated more concisely, the time-temperature superposition principle says that E(T,t) = E(Tr, tr) (46)

The dependence of the shift factor, aT, on temperature is given by Williams-Landel-Ferry (WLF) relationship log aT

(47)

Where C1 and C2 are constants for a given polymer and Tr is the reference temperature. It is common practice now to use the glass transition temperature measured by a very slow rate method as the reference temperature for master curve construction. Then the shift factor for most amorphous polymers is given fairly well by log10 aT =

(48)

The constants C1 and C2 depend on the material. "Universal" values are C1 = 17.4 and C2 = 51.6; however, significant deviations from these values may exist for some polymers. If C1 and C2 are not known, each curve may be horizontally shifted to the reference temperature curve until a value of modulus on the shifted curve coincides with one on the reference curve. The shift factor may then be calculated by use of the defining relationship for aT given by eq. 18. The WLF parameters, C1 and C2, may then be determined by plotting (T Tr) / log aT versus (T Tr), where rearrangement of eq. 20 gives The procedure of horizontal shifting is illustrated below: The relaxations that lead to viscoelastic behaviour are the result of various types of molecular motions, and they occur more rapidly at higher temperature. The compliance D(t), for instance, is therefore a function of temperature, T, which means that it should really be written as D(t,T). Suppose that the effect of a rise in temperature from some chosen standard temperature To is to speed up every stage in a relaxation process by a constant factor that depends on the new temperature T. This is equivalent to saying that the interval of time required for any small change in strain to take place is divided by a factor aT that depends on T and has the value 1 when T = To. This means that, if measured values of D(t,T) are plotted against taT, curves for all temperatures should be superposed. Because the time over which the compliance changes are very large, it is usual to use logarithmic scales for t. When this is done it is necessary to shift the curves for different temperatures by a constant amount log aT or log aT along the t axis in order to get superposition with the curve for To. The quantity aT is therefore called the shift factor. Figure 5 shows data for a set of measurements before and after the shifts have been applied. The idea that the same effect can be produced on the compliances or moduli either by a change of temperature or by a change of time-scale is called timetemperature equivalence. (49)

Fig. 11 The curve obtained after the shifts have been applied is called a master curve. It is important to realise, however, that a master curve is obtained only if there is just one important relaxation process in the effective range of temperatures studied. This is because there is no reason why the shift factors for two different relaxation processes should be the same at any temperature. Timetemperature superposition is, however, observed over a wide range of frequencies for many non-crystalline polymers in which the glass transition is the only important relaxation process. The left-hand panel of Fig. 12 contains sketches of typical stress relaxation curves for an amorphous polymer at a fixed initial strain and a series of temperatures. Such data can be obtained much more conveniently than those where the modulus was measured at a given time and a series of temperatures. It is found that the stress relaxation curves can be caused to coincide by shifting them along the time axis. This is shown in the right-hand panel of Fig. 12 where all the curves except that for temperature Tg have been shifted horizontally to form a continuous "master curve" at temperature Tg. The glass transition temperature is shown here to be T5 at a time of 10-2 min. The polymer behaves in a glassy manner at this temperature when a strain is imposed within 10-2 min or less.

Fig. 12. Left panel: Stress decay at various temperatures T1 < T2 <.. T9. Right panel: Master curve for stress decay at temperature T8. Similar curves can be constructed for creep test data of amorphous polymers. Typical creep curves of a low Tg polymer measured over four decades of time and at several temperatures are shown in fig.13, As illustrated creep compliance increases with increasing temperature. This means that the sample sotens as

temperature is increased.. Individual creep curves obtained at different temperatures over short time periods (Fig. 13A) can be shifted horizontally to yield a master curve at a given temperature and covering a wide range of temperatures. Such a creep curve showing three regions of viscoelastic behavior is shown in Fig. 13B. Fig. 13 A. Plots of creep compliance as a function of time at different temperatures. Arrows show direction of the horizontal shift of data to obtain a master curve at a reference temperature taken as the Tg of the polymer. B. Creep master curve obtained by timetemperature superposition. A

A complete picture of the behavior of the material is obtained in principle by operating in experimentally accessible time scales and varying temperatures. Time-temperature superposition can be expressed mathematically as E(T1,t) = E(T2, t/aT) (50)

for a stress relaxation experiment. The effect of changing the temperature is the same as multiplying the time scale by shift factor aT. Accumulation of long-term data for design with plastics can be very inconvenient and expensive. The equivalence of time and temperature allows information about mechanical behavior at one temperature to be extended to longer times by using data from shorter time studies at higher temperature. It should be used with caution, however, because the increase of temperature may promote changes in the material, such as crystallization or relaxation of fabrication stresses, that affect mechanical behavior in an irreversible and unexpected manner. Note also that the master curve in the previous figure is a semilog representation. Data such as those in the left-hand panel are usually readily shifted into a common relation but it is not always easy to recover accurate stress level values from the master curve when the time scale is so compressed. The following simple calculation illustrates the very significant temperature and time dependence of viscoelastic properties of polymers. It serves as a convenient, but less accurate, substitute for the accumulation of the large amount of data needed for generation of master curves. Suppose that a value is needed for the compliance (or modulus) of a plastic article for 10 years service at 25C. What measurement time at 80C will produce an equivalent figure? We rely here on the use of a shift factor, aT, and Eq. (23). Assume that the temperature dependence of the shift factor can be approximated by an Arrhenius expression of the form aT = exp H/R [1/T 1/T0] (51)

where the activation energy, H, may be taken as 0.12 MJ/mol, which is a typical value for relaxations in semicrystalline polymers and in glassy polymers at temperatures below T g. (The shift factor could also have been calculated from the WLF relation if the temperatures had been around Tg of the polymer.) In the present case: aT = exp[0.12 x 10-6 / 8.31] [ 1/353 1/298] = 0.53 X 10 -3 (52) The measurement time required at 80C is [0.53 x 10 -3] [10 years] [365 days/year] [24 hours/day] = 45.3 hours, to approximate 10 years service at 25C. The relaxation time constant is proportional to the viscosity of the polymer, and therefore, decrease exponentially with an increase in temperature. Since relaxation time is the reciprocal of a rate, we can relate it to the temperature in Kelvin T by an Arrhenius type rate equation as 1/ C Q T = = = = C e-Q/RT (53)

rate constant independent of temperature activation energy for the process temperature in Kelvin and R = gas constant (8.314 J/mol.K)

The viscoelastic behavior of polymeric material is dependent on both time and temperature; several experimental techniques may be used to measure and quantify this behavior. Stress relaxation measurements represent one possibility. In a stress relaxation test a specimen is quickly loaded at (T Tg) to a constant stress level, 0, that is maintained throughout the test. The stress is measured as a function of time, while keeping the temperature constant. Stress is observed to decay with time due to molecular relaxation process that takes place within the polymer. We may define a relaxation modulus Er(t) a time dependent elastic modulus as Er(t) = (t) / 0 (t) is the measured timedependent stress and 0 is the strain level, which is maintained constant. It shows that the modulus also decreases exponentially with time during a stress relaxation experiment. Furthermore, the magnitude of the relaxation modulus is a function of temperature; and to more fully characterize the viscoelastic behavior of a polymer, isothermal stress relaxation measurement must be constructed over a range of temperatures. Fig. 10 is a schematic log Er(t) v/s log time plot for a polymer that exhibits a visocelastic behavior; included are several curves generated at different temperatures. Note that (1) the decrease of Er(t) with time (i.e. stress decay with time) and (2) the displacement of the curves to lower Er(t) levels with increasing temperatures.

Fig. 14 To represent the influence of temperature, data points are taken at a specific time from the log E r(t) versus log time plot e.g. t1 (Fig 14) and then cross plotted as log Er(t) versus temperature. Fig 15 is such a plot for an amorphous (atactic) polystyrene, t1 in this case was arbitrarily taken 10 s after the load application.

Figure 15 Several distinct regimes may be noted on the curve shown in this figure. For the first, at the lowest temperatures, in the glassy region, the material is rigid and brittle, and the value of Er (10) is that of the elastic modulus, which initially is virtually independent of temperature, over this temperature range, the strain time characteristics are as represented in Fig 2b. On a molecular level, the long molecular chains are essentially frozen in position at these temperatures. As the temperature is increased, Er (10) abruptly drops by about a factor of 103 within a 200C temperature span; this is sometimes called the leathery, or glass transition region, and T g lies near the upper temperature extremity; for polystryrene (Fig 15, Tg = 100oC). Within this temperature region a polymer specimen will be leathery, that is, deformation will be time dependent and not totally recoverable on release of an applied load, as shown in Fig. 2c. Over these temperatures, the atomic vibrations are such that the molecules begin to experience coordinated chain motions.

Within the rubbery plateau temperature region, the material deforms in a rubbery manner; here, both elastic and viscous components are present, and deformation is easy to produce because the relaxation modulus is relatively low. The final two high temperature regions are rubbery flow and viscous flow. Upon heating through these temperatures, the material experiences a gradual transition to a soft rubbery state, and finally to a viscous liquid. Within the viscous flow region, the modulus decreases dramatically with increasing temperatures and again the strain time behavior is as represented in Fig. 2d. From a molecular standpoint, chain motion intensifies so greatly that for viscous flow, the chain segments experience vibration and rotational motion quite independently of one another. At these temperatures, any deformation is entirely viscous.

1. 2. 3. 4.

Glassy state: Frozen molecules No rotational or transitional motion Leathery region Some rotational motion No translational motion Rubbery region Large degree of rotational motion Small degree of translational motion Viscous region Complete rotational & translational motion of the whole polymer molecule

Normally, the deformation behavior of a viscous polymer is specified in terms of viscosity, a measure of materials resistance to flow by shear forces. The rate of stress application also influences the viscoelastic characteristics. Increasing the loading rate has the same influence as lowering temperature. The log Er(t) versus temperature behavior for polystyrene materials having several molecular configurations is plotted in Figure 12. The curve for the amorphous material (curve C) is the same as in Figure 2. For a lightly crosslinked atactic polystyrene (curve B), the rubbery region forms a plateau that extends to the temperature at which the polymer decomposes; this material will not experience melting. For increased crosslinking, the magnitude of the plateau Er(10) value will also increase. Rubber or elastomeric materials display this type of behavior and are ordinarily used at temperatures within this plateau range.

Figure 12 Logarithm of the relaxation modulus versus temperature for amorphous polystyrene, showing the five different regions of viscoelastic behavior. Also shown in Figure 12 is the temperature dependence for an almost totally crystalline isotactic polystyrene (curve A).The decrease in at is much less pronounced than the other polystyrene materials since only a small volume fraction of this material is amorphous and experiences the glass transition. Furthermore, the relaxation modulus is maintained at a relatively high value with increasing temperature until its melting temperature is approached. From Figure 12, the melting temperature of this isotactic polystyrene is about 240C. Viscoelastic Creep Many polymeric materials subjected to a load may creep (time dependent deformation). That is, their deformation under a constant applied load at a constant temperature continues to increase with time. The magnitude of the strain increment increases with increased applied stress and temperature. This type of deformation may be significant even at room temperature an under modest stresses which lie below the yield strength of the material. For example, automobile tyres may develop flat spots on their contact surfaces when the automobile is parked for a long period. In creep test, a stress (normally tensile) is applied instantaneously which is maintained at a constant level while strain is measured as a function of time. Tests are performed under isothermal conditions.

a) Applied stress and b) induced strain (b) as functions of time over a short period for a viscoelastic material.

The corresponding relaxation modulus or creep modulus (time dependent) expression is Ec (t) = (0) / (t) (0) = constant applied stress

(t)

time dependent strain.

The creep modulus is also temperature sensitive and diminishes with increasing temperature. The temperature at which the creep of polymeric material takes place is also an important factor determining the creep rate. At temperature below the glass transition temperature for thermoplastics, the creep rate is relatively low due to restricted molecular chain mobility. Above their Tg, TP deform easily by a combination of elastic and plastic deformation which is referred as viscoelastic behavior. Above Tg, the molecular chain slide past each other more easily, and so this type of easier deformation is sometimes referred to as viscous flow.

Вам также может понравиться