Published Date
International Journal of Solids and Structures
1 June 2009, Vol.46(11):2298–2308, doi:10.1016/j.ijsolstr.2009.01.015
Open Archive, Elsevier user license
Author
Keywords
Thermoplastic elastomers
Polymer composite
Viscoplasticity
Finite strains
Constitutive equations
For further details log on website :
http://www.sciencedirect.com/science/article/pii/S0020768309000419
International Journal of Solids and Structures
1 June 2009, Vol.46(11):2298–2308, doi:10.1016/j.ijsolstr.2009.01.015
Open Archive, Elsevier user license
Author
Received 28 March 2008. Revised 15 January 2009. Available online 28 January 2009.
Abstract
Observations are reported on carbon black–filled thermoplastic elastomer (TPE) in uniaxial loading–unloading tensile tests with various strain rates (ranging from to ) at temperatures in the interval from 25 to 90 °C. A constitutive model is derived for the viscoplastic response of a TPE composite at three-dimensional deformations with finite strains. The stress–strain relations involve six adjustable parameters that are found by fitting the experimental data. It is shown that the model correctly describes the observations, and its parameters are affected by temperature and strain rate in a physically plausible way.
1 Introduction
This paper is concerned with the experimental and theoretical analysis of the mechanical response of a carbon black-reinforced thermoplastic elastomer (TPE) at quasi-static deformations with finite strains.
Thermoplastic elastomers are an important class of polymers that combine mechanical properties of rubbers with high-speed processability and recyclability of thermoplastics (Holden et al., 2004 and Grady and Cooper, 2005). The experimental investigation focuses on the viscoplastic behavior of Thermoplast K reinforced with a relatively large amount (about 60–70 wt.%) of carbon black particles. The polymeric matrix of this composite consists of a hydrogenated styrene block copolymer (HSBC)-based thermoplastic elastomer with addition of polypropylene segments. A rather complicated chemical structure of the TPE composite and high concentration of filler ensures its strong thermal and chemical resistance. Due to these properties, Thermoplast K may be employed as a sealing material for low-temperature proton exchange membrane fuel cells (PEM FC) with operating temperatures up to 100 °C. From the standpoint of applications, the objectives of this work are (i) to study the mechanical response of Thermoplast K at various temperatures, and (ii) to develop a simple constitutive model for its thermo-viscoplastic response to be used in finite-element analysis of a fuel cell.
Mechanical properties of thermoplastic elastomers has attracted substantial attention in the past decade (Boyce et al., 2001a, Boyce et al., 2001b, Boyce et al., 2001c, Blundell et al., 2002, Blundell et al., 2004, Le et al., 2003, Krijgsman and Gaymans, 2004, Baeurle et al., 2005, Indukuri and Lesser, 2005, Qi and Boyce, 2005, Drozdov, 2006, Drozdov, 2007, Drozdov and Christiansen, 2006, Drozdov and Christiansen, 2007, Laity et al., 2006, Long and Sotta, 2006, Yi et al., 2006, Roland et al., 2007, Sarva et al., 2007 and Eceiza et al., 2008). Most of these studies, however, concentrated on the time- and rate-dependent behavior of unfilled TPEs at ambient temperature. Only a few works dealt with the viscoelastic and viscoplastic responses of thermoplastic elastomers reinforced with carbon black (Kurian et al., 1995, Datta et al., 1996, Katbab et al., 2000 and Yamauchi et al., 2005), and none of them investigated mechanical properties at elevated temperatures. From the viewpoint of fundamental research, the objectives of the present study are (i) to report observations in uniaxial tensile tests with finite strains on a carbon black-reinforced TPE in a relatively large range of temperatures, and (ii) to analyze the effects of temperature and strain rate on adjustable parameters in the stress–strain relations.
Derivation of the constitutive model is based on a homogenization concept, according to which a complicated morphology of a TPE composite is replaced with a simple structure that captures essential features of its mechanical response. Following common practice (Tomita et al., 2007), a thermoplastic elastomer is thought of as a permanent non-affine network of chains bridged by junctions. The assumption that the network is permanent means that detachment of active chains from their junctions and merging of dangling chains with the network are neglected. With reference to Green and Tobolsky, 1946 and Tanaka and Edwards, 1992, the latter is tantamount to the neglect of viscoelasticity of the equivalent network. Non-affinity presumes that junctions slide with respect to their reference positions under deformation, which implies that the viscoplastic response is associated with sliding (plastic flow) of junctions. To simplify the analysis, it is postulated that plastic flow is volume-preserving, the strain energy of a chain is described by the neo-Hookean formula, and the energy of interaction between chains is accounted for by the incompressibility condition for the equivalent network.
The exposition is organized as follows. Observations in uniaxial tensile loading–unloading tests and relaxation tests at various temperatures are reported in Section 2. A constitutive model in finite viscoplasticity of TPE composites is developed in Section 3. In Section 4, the stress–strain relations are simplified for uniaxial tension of an incompressible medium. Adjustable parameters in the governing equations are found in Section 5 by fitting the experimental data. Results of numerical simulation for simple shear of a TPE composite are reported in Section 6. Some concluding remarks are formulated in Section 7.
2 Experimental results
Carbon black-reinforced thermoplastic elastomer Thermoplast K TV5LVZ [density , melt flow index 12 g/10 min at 230 °C, elongation at break 520%, hardness (shore A) 50] was purchased from Kraiburg TPE GmgH (Germany). Dumbbell specimens for uniaxial tensile tests (ASTM standard D638) with cross-sectional area 10.2 mm × 3.4 mm were molded by using injection-molding machine Arburg 320C.
Mechanical tests were conducted with the help of universal testing machine Instron-5569 equipped with a thermal chamber and an electro-mechanical sensor for control of longitudinal strains in the active zone of samples. The tensile force was measured by a standard load cell. The engineering stress σ was determined as the ratio of the axial force to the cross-sectional area of specimens in the stress-free state.
The experimental program involved two series of tests. The first consisted of tensile loading–unloading tests (1 cycle) with the maximum engineering strain conducted at the temperatures T = 25, 50, 70, and 90 °C with cross-head speeds of 1, 5, 10, 20, 100, and 150 mm/min. These cross-head speeds corresponded to the strain rates of , , , , , and , respectively. In a test, a sample was stretched up to the maximum strain ϵmax with a constant strain rate and unloaded down to the zero stress with the strain rate . Each test was performed on a new specimen. In all tests at elevated temperatures, samples were pre-heated in a thermal chamber at the required temperature for at least half an hour, and, afterwards, equilibrated in the testing machine for 10 min before deformation.
To assess repeatability of observations, several tests were carried out on three different specimens (appropriate data are not presented). The maximum difference between the engineering stresses measured on different samples did not exceed 4%. This distinguishes the mechanical response of carbon black-reinforced TPEs from that of filled rubbers, as cyclic preloading of specimens is necessary in the latter case to reach an acceptable level of repeatability of measurements (Haupt and Sedlan, 2001).
The range of temperatures was chosen to cover the entire interval of operating temperatures of a low-temperature PEM FC. The interval of strain rates was restricted by the requirement that the duration of a test did not exceed 40 min (from below) and by the maximal cross-head speed of the testing machine (from above). The maximum engineering strain in tensile tests ϵmax was limited by the range of admissible strains for the extensometer (less than 100%).
Observations in cyclic tensile tests at various temperatures T are reported in Fig. 1, Fig. 2, Fig. 3 and Fig. 4, where the engineering tensile stress σ is plotted versus tensile strain ϵ. The following conclusions are drawn:
- 1.The stress–strain diagrams at loading are strongly affected by cross-head speed. The engineering tensile stress grows with strain rate . The maximum increase in stress is about 27% at room temperature, 35% at T = 50, 40% at T = 70, and 65% at T = 90 °C, which means that the effect of strain rate on σ becomes more pronounced at elevated temperatures.
- 2.At room temperature, the stress–strain diagrams at unloading are practically independent of cross-head speed. With the growth of temperature, the influence of strain rate on σ at retraction becomes stronger. At the maximum temperature T = 90 °C, the tensile stress at unloading noticeably increases with .
- 3.The residual engineering strain ϵ (the strain at which the tensile stress vanishes at unloading) is practically independent of strain rate and temperature. In all tests, this strain belongs to the interval between 10 and 15%.
- 4.Given a strain rate , the engineering tensile stress is strongly affected by temperature. At the highest strain rate of , the maximum tensile stress (reached at the maximum strain ϵmax) equals 1.50, 1.13, 0.68, and 0.38 MPa at the temperatures T = 25, 50, 70, and 90 °C, respectively. This means that the maximum stress decreases by a factor of 4 with growth of temperature in the interval under consideration.
The other series of tests involved uniaxial tensile relaxation tests with the engineering strain conducted at the temperatures T = 25, 50, 70, 90, and 110 °C. In each test, a specimen was loaded with a constant cross-head speed of 100 mm/min up to the required strain ϵ. Afterwards, a decrease in stress was measured as a function of time while the strain was preserved constant. Following the ASTM protocol for short-term relaxation tests (ASTM STP 676), the duration of relaxation tests was chosen.
The experimental data in relaxation tests are reported in Fig. 5, where the engineering stress σ is plotted versus relaxation time , where is the instant when the relaxation process starts. Following common practice, the semi-logarithmic plots are used with . The following conclusions are drawn:
- 1.The relaxation curves are typical of elastomers. They demonstrate a practically linear decay in tensile stress with logarithm of relaxation time.
- 2.The tensile stress strongly decreases with temperature at all relaxation times . The slopes of the curves decay with temperature.
- 3.The amounts of stress relaxing within the time weakly decrease with temperature T and equal 34.6, 33.5, 31.8, 28.3, and 22.8% at the temperatures T = 25, 50, 70, 90, and 110 °C, respectively.
3 Constitutive model
With reference to the homogenization concept (Bergstrom et al., 2002), a carbon black-filled thermoplastic elastomer (polymer composite with a complicated micro-structure) is thought of as an equivalent one-phase continuum. A non-affine network of flexible chains bridged by permanent junctions is chosen as the equivalent structure. To simplify derivations, we presume macro-deformation of the equivalent network to be volume-preserving. The incompressibility condition is in good agreement with available experimental data for the Poisson ratio (close to 0.5) of thermoplastic elastomers (Kojima et al., 2004).
3.1 Kinematic equations for sliding of junctions
Under deformation, junctions in a non-affine polymer network slide with respect to their reference positions. Denote by the deformation gradient for macro-deformation and by the deformation gradient for sliding (plastic flow) of junctions. According to the multiplicative decomposition of the deformation gradient, the deformation gradient for elastic deformation reads
1
where the dot stands for inner product. The incompressibility conditions for macro-deformation and plastic flow together with Eq. (1) imply that the elastic deformation is volume-preserving.
Differentiating Eq. (1) with respect to time and introducing the velocity gradients by the conventional formulas
2
we find that
3
where the velocity gradient for elastic deformation is given by
4
The left and right Cauchy–Green tensors for elastic deformation are determined by
5
where stands for transpose. It follows from the incompressibility condition for elastic deformation that the third principal invariants of these tensors equal unity, whereas the first and second invariants read
6
where the colon stands for convolution, and I is the unit tensor.
7
where
8
is the rate-of-strain tensor and
9
is the vorticity tensor for elastic deformation. Combination of Eqs. (5), (6) and (7)results in the formula
10
To describe plastic flow of junctions, it is postulated that
- 1.The rate-of-strain tensor for elastic deformation is connected with the rate-of-strain tensor for macro-deformation11by the linear relation12where ϕ is a scalar function to be determined in what follows,
- 2.The vorticity tensor for elastic deformation vanishes13
14
15
16
To deduce another form of Eq. (16), we introduce the left and right Cauchy–Green tensors for macro-deformation
17
By analogy with Eq. (7), we find that
18
It follows from Eq. (2) that
Insertion of these expressions into Eq. (16) results in the differential equation
19
which expresses the deformation gradient for sliding of junctions by means of the deformation gradient for macro-deformation.
The approach based on Eqs. (12) and (13) slightly differs from the standard hypotheses in finite viscoplasticity of solid polymers. As there is no physically reasonable way to distinguish between elastic and plastic vorticity tensors, the conventional theories (i) postulate that the vorticity tensor for plastic deformation vanishes [which allows the vorticity tensor for elastic deformation to be found from Eqs. (4) and (9)] and (ii) introduce algebraic relations between the rate-of-strain tensor for plastic deformation and the rate-of-strain tensor for macro-deformation and the Cauchy stress tensor Σ. Derivation of the latter equations leads, however, to some complications, as the tensor on the one hand and the tensors and Σ on the other are defined in different bases.
Within the present approach, kinematic equations are formulated for components of the velocity gradient for elastic deformations . These tensors are defined in the same basis as the tensor , which implies that Eq. (12) is objective. It follows from Eqs. (12) and (13) that the deformation gradient can be found from Eq. (19), which means that plastic flow in the equivalent network is entirely determined by history of macro-deformation.
3.2 Stress–strain relations
The strain energy of a polymer chain is determined by the conventional formula for a neo-Hookean medium
20
where μ stands for rigidity per chain. The strain energy per unit volume of a network equals the sum of strain energies of individual chains
21
where , and N stands for the number of chains per unit volume. Eq. (21) is grounded on the assumption that the energy of inter-chain interactions is accounted for by means of the incompressibility condition only.
Eq. (20) coincides with an expression for the strain energy of a Gaussian chain developed within the statistical theory of rubber elasticity (in the latter case, the coefficient is proportional to absolute temperature T). It is worth noting that Eq. (20)is adopted for its simplicity only, which means that no proportionality is presumed between rigidity of a chain and temperature of the network. Although a decrease in the elastic modulus μ with T may be explained within the concept of polymer networks with constrained junctions (Drozdov et al., 2008), we do not dwell on this issue in the present study.
22
The Clausius–Duhem inequality for isothermal deformation of an incompressible medium reads
23
where Q stands for internal dissipation per unit volume and unit time, and denotes the deviatoric component of the Cauchy stress tensor Σ. Inserting Eq. (22) into Eq. (23) and assuming internal dissipation to vanish, we arrive at the stress–strain relation
24
where p denotes an unknown pressure.
3.3 Kinetics of plastic flow
To describe changes in the coefficient ϕ in Eq. (24) at active loading and unloading, we presume evolution of ϕ with time to be governed by strain energy of the equivalent network W. The growth of ϕ with time at active loading is determined by the differential equation
25
where are adjustable parameters. Eq. (25) together with the initial condition implies that the function monotonically increases with time and reaches its ultimate value Φ at relatively large deformations. With reference to Eq. (12), this means that the rate-of-strain tensor for elastic deformation coincides with the rate-of-strain tensor for macro-deformation at the initial instant (no sliding of junctions in the reference state), and it becomes proportional to (with the coefficient of proportionality ) at the steady regime of plastic flow. According to Eq. (21), the rate of increase in ϕ is proportional to the dimensionless strain energy , whereas the coefficient of proportionality decays due to sliding of junctions and vanishes when ϕ equals its maximum value Φ.
Eq. (25) may be treated as a kinetic equation for a chemical reaction of the second order. Although some justification for this order can be presented (sliding of junctions in the polymer matrix induces sliding of aggregates of filler, which, in turn, accelerates plastic flow in the rubbery phase), Eq. (25) is treated as a merely phenomenological relation.
Evolution of the coefficient ϕ at unloading is described by the differential equation similar to Eq. (25),
26
where , , and are adjustable parameters. The initial condition for Eq. (26) expresses continuity of ϕ at the instant when transition occurs from active loading to unloading.
Eq. (26) implies that the function ϕ increases with time at unloading. The first term in the right-hand side of Eq. (26) coincides formally with an appropriate term in Eq. (25), whereas the last term characterizes self-acceleration of plastic flow at retraction. This term is absent in Eq. (25) for deformation of a virgin material, but a similar term (with a negative coefficient) should appear in an analog of Eq. (25) for reloading of a thermoplastic elastomer under cyclic deformation. We do not dwell on this refinement as the present study deals with the first cycle of deformation only.
3.4 Adjustable parameters
Eqs. (24), (25) and (26) provide a set of stress–strain relations for an arbitrary three-dimensional deformation of a particulate composite with TPE matrix at finite strains. These equations involve six adjustable parameters:
- 1.The coefficient μ characterizes elasticity of the equivalent network.
- 2.The coefficients a, α, and β describe evolution of the rate of plastic flow at active loading and unloading.
- 3.The quantities determine the rate of sliding of junctions in the steady regime of plastic flow under active deformation and retraction.
In general, these quantities depend on strain rate and temperature. To avoid formulation of an overly complicated constitutive model with 6 unknown functions of two variables, the following simplifications are introduced:
- (A)The elastic modulus μ is independent of strain rate. The effect of temperature T on μ is described by the linear equation27where are positive coefficients.
- (B)The steady rate of plastic flow at loading Φ is independent of temperature T and weakly (logarithmically) decreases with strain rate28where are positive coefficients, and29stands for the equivalent strain rate for macro-deformation.
- (C)The steady rate of plastic flow at unloading Ψ is independent of strain rate, and linearly decreases with temperature,30where are positive coefficients.
- (D)The coefficients a, α, and β are independent of temperature and proportional to the equivalent strain rate31where , , and are positive constants.
4 Uniaxial deformation
Uniaxial tension of an incompressible specimen is described by the formulas
32
where are Cartesian coordinates in the initial and actual states, respectively, and stands for elongation ratio. The deformation gradient for macro-deformation reads
33
where are basic vector of the Cartesian frame in the refeence state. Insertion of Eq. (33) into Eqs. (2) and (11) implies that
34
where
35
Transition from the initial state into the intermediate (stress-free) state is described by the equations similar Eq. (32),
where are Cartesian coordinates in the stress-free state, and is a function to be found. By analogy with Eq. (33), one can write
36
37
38
According to Eq. (38), Eq. (13) is satisfied identically. Insertion of Eqs. (34) and (38)into Eq. (12) implies the differential equation
39
where
Excluding the unknown pressure from the boundary condition and introducing the engineering tensile stress , we arrive at the formula
40
41
describe uniaxial deformation of a TPE composite at tension–compression.
5 Fitting of observations
Adjustable parameters in the governing equations are found by fitting the observations reported in Fig. 1, Fig. 2, Fig. 3 and Fig. 4. Each set of experimental data is matched separately.
5.1 Loading
We begin with approximation of loading paths of the stress–strain curves, when the elongation ratio k obeys the differential equation
To determine the quantities μ, a and Φ, we fix some intervals , where the best-fit parameters a and Φ are assumed to be located, and divide these intervals into sub-intervals by the points with , . For each pair , the stress–strain relations are integrated numerically from to with . The governing equations are presented in the form
42
Integration of Eq. (42) is performed by the Runge–Kutta method with the step . The modulus μ is found by the least-squares technique from the condition of minimum of the function
43
where summation is performed over all elongation ratios at which observations are reported, is the engineering tensile stress measured in a test, and is given by Eq. (40). The best-fit quantities a and Φ are found from the condition of minimum of function (43). Afterwards, the initial intervals are replaced with the new intervals , , and the calculations are repeated three times. The best-fit material parameters μ, a, and Φ are plotted versus strain rate in Fig. 6, Fig. 7and Fig. 8.
In accord with assumption (A), the dependencies of μ on (found by matching observations at each temperature T separately) are approximated in Fig. 6 by constants. These constants are plotted versus temperature T in Fig. 12, where the experimental data are matched by Eq. (27). The coefficients in Eq. (27)are calculated by the least-squares method and are listed in Table 1.
Table 1. Adjustable parameters for TPE composite.
Parameter | Dimension | Value |
---|---|---|
MPa | 2.19 | |
MPa K−1 | ||
s−1 | ||
s−1 | ||
s−1 | ||
K−1 |
Following assumption (B), the dependencies of Φ on (determined at all temperatures T under consideration) are approximated by Eq. (28), where is replaced with and the coefficients are found by the least-squares method. These coefficients are collected in Table 1.
With reference to assumption (D), the dependencies of a on (found at all temperatures T) are approximated by the equation
where the coefficient is calculated by means of the least-squares algorithm (its best-fit value is reported in Table 1).
5.2 Unloading
We proceed with fitting unloading paths of the stress–strain diagrams when the elongation ratio k is governed by the differential equation
For each temperature and strain rate, Eq. (42) are first integrated with the adjustable parameters collected in Table 1. Afterwards, we fix some intervals , , and , where the parameters α, β, and Ψ are assumed to be located, and divide these intervals into sub-intervals by the points , , and with , , . For each triplet , the stress–strain relations are integrated numerically from to . The governing equations are presented in the form
44
Integration is carried out by the Runge–Kutta method with the step . The best-fit parameters α, β, and Ψ are determined from the condition of minimum of function (43), where is given by Eq. (40). When these quantities are found, the initial intervals are replaced with the new intervals , , , and the calculations are repeated three times.
The best-fit material parameters α, β, and Ψ are depicted versus strain rate in Fig. 9, Fig. 10 and Fig. 11. With reference to assumption (D), the experimental dependencies are approximated by the relations
45
where are determined by the least-squares technique. Eq. (45) coincide with Eq. (31), where is replaced with . The best-fit values of are reported in Table 1.
Fig. 11 shows that the parameter Ψ is practically independent of strain rate, in accord with hypothesis (C). The experimental data depicted in this figure are approximated by constants. The dependence of these constants on temperature is illustrated in Fig. 12, where the data are matched by Eq. (30). The coefficients in Eq. (30)are determined by the least-squares algorithm and are listed in Table 1.
5.3 Discussion
Fig. 1, Fig. 2, Fig. 3 and Fig. 4 demonstrate good agreement between the observations in tensile loading–unloading tests at various temperatures and the results of numerical simulation.
Despite some scatter of the data depicted in Fig. 5 and Fig. 11, these figures confirm that the parameters are independent of strain rate. Fig. 12 shows that the effect of temperature on their average values (over tests with various strain rates at fixed temperatures) is adequately described by Eqs. (27) and (30).
Although Fig. 7 and Fig. 8 reveal an acceptable agreement between the experimental data and their approximations by Eqs. (28) and (31), the scatter of observations is noticeable. It may be explained by the fact that the loading paths of the stress–strain diagrams are rather simple curves that do not require three adjustable parameters for their fitting. Deviations of the data depicted in Fig. 9 and Fig. 10 from their approximations by Eq. (45) are noticeably lower, which means that kinetic Eq. (26)provides an adequate model for viscoplastic flow of junctions at unloading.
6 Numerical simulation
To assess ability of the constitutive equations to predict mechanical properties of TPE composites, we concentrate of their thermo-viscoplastic response at simple shear. The importance of our analysis for applications is grounded on the assumption that shear provides the main mode of deformation of sealing materials in fuel cells (Wang and Wang, 2005 and Malzbender et al., 2007). From the standpoint of fundamental research, the study of simple shear allows basic hypotheses of the model to be examined for macro-deformations with non-vanishing vorticity tensors.
Simple shear of an infinite layer is described by the equations
where stands for shear. The deformation gradient for macro-deformation reads
46
47
48
Combination of this relation with Eq. (14) results in the differential equation
49
The initial condition for Eq. (49) reads . We search the tensor in the form
50
where are unknown functions of time that obey the incompressibility condition
51
52
It follows from the third equality in Eq. (52) that
53
The first two equalities in Eq. (52) imply that the functions coincide. This conclusion together with Eqs. (51) and (52) yields
54
55
where the shear stress Σ reads
56
It follows from Eqs. (25), (26), (31), (48) and (55) that the constitutive equations for active loading with a constant shear rate read
57
and the governing equations for unloading with the shear rate are given by
58
To assess intensity of stresses at cyclic loading of the TPE composite, numerical simulation is performed of Eqs. (56), (57) and (58) for simple shear with the strain rates and the maximum shears , 2.0, and 2.5 at the temperatures T = 60, 80, and 100 °C. Integration of the governing equations is carried out by the Runge–Kutta method with the step . The stress Σ is plotted versus shear k in Fig. 13, Fig. 14 and Fig. 15. The following conclusions are drawn:
- 1.The shear stress Σ monotonically increases with k at active loading and decreases at unloading. The functions are strongly nonlinear both at loading and unloading.
- 2.The stress–strain diagrams are substantially affected by temperature. The maximum stress noticeably decreases with T. For example, the shear stress Σ at equals 4.31, 2.96, and 1.56 MPa at the temperatures T = 60, 80, and 100 °C, respectively.
- 3.Given a temperature T, the shear stress increases with strain rate, but this growth is relatively weak. When the strain rate increases by 5 orders of magnitude, the maximum stress (at ) grows by 90% at all temperatures under consideration.
- 4.The residual shear k (at which Σ vanishes under retraction) strongly increases with maximum shear . This quantity is practically independent of temperature, and it equals 0.53, 1.26, and 1.90 for cyclic deformation with the strain rate , and 0.62, 1.34, and 1.98 for loading–unloading with the strain rate .
Although the above results seem plausible, they should be treated with caution for two reasons: (i) observations were obtained in uniaxial tensile tests with the maximum elongation ratio , whereas numerical simulation was performed for shear tests with the larger maximum shear , and (ii) adjustable parameters for unloading α, β, and Ψ may, in general, depend on , but the effect of maximum deformation on these quantities was not investigated experimentally.
7 Concluding remarks
Observations have been reported in uniaxial tensile tests on carbon black–filled thermoplastic elastomer Thermoplast K at various temperatures in the range from ambient temperature to 90 °C. The experimental data reveal an interesting phenomenon: (i) the unloading curves measured at various strain rates practically coincide at room temperature, (ii) the difference between these curves grows with temperature, and (iii) it becomes significant at the highest temperature under consideration.
Constitutive equations have been developed for the viscoplastic response of TPE composites. With reference to the homogenization concept, a composite is treated as an equivalent non-affine network of flexible chains linked by permanent junctions. Non-affinity means that macro-deformation induces sliding of junctions with respect to their reference positions. The effect of strain on the rate of sliding is described by Eqs. (25) and (26).
Modeling a carbon black-reinforced composite as an equivalent one-phase continuum noticeably simplifies derivations and allows the number of adjustable parameters to be reduced substantially. It leads, however, to the neglect of some effects driven by deformation-induced evolution in distribution of filler (Klüppel, 2003). In particular, this approach disregards mechanically induced anisotropy of filled elastomers under cyclic loading. The latter can be described by means of the constitutive models developed by Ihlemann, 2005, Diani et al., 2006a, Diani et al., 2006b and Itskov et al., 2006.
Stress–strain relations for an arbitrary three-dimensional deformation with finite strains are derived by using the laws of thermodynamics. These equations involve 6 material parameters that are found by matching experimental data in tensile loading–unloading tests with various strain rates. Good agreement is demonstrated between the observations and the results of numerical simulation. An advantage of the model is that loading and unloading paths of the stress–strain diagrams are described by similar governing equations (25) and (26).
Phenomenological equations (27), (28), (30) and (31) are proposed to describe the effect of temperature and strain rate on the adjustable parameters. Fig. 6, Fig. 7, Fig. 8, Fig. 9, Fig. 10, Fig. 11 and Fig. 12 reveal that these relations are satisfied with an acceptable level of accuracy.
The constitutive equations have been applied to predict the effect of temperature and strain rate on the shear stress at simple shear of a TPE composite. The results of numerical simulation show that the shear stress noticeably grows with strain rate and decreases with temperature. It is revealed that the residual strain after unloading to the zero stress strongly increases with maximum strain under loading. The curves presented in Fig. 13, Fig. 14 and Fig. 15 appear to be physically plausible, which may serve as a justification of assumptions (12) and (13) that slightly differ from the conventional hypotheses in finite viscoplasticity of solid polymers.
As our purpose is to develop simple stress–strain relations to be employed in finite-element analysis, a number of assumptions have been made in the derivation of constitutive equations. The most important among them are that (i) deformation of the equivalent polymer network is volume-preserving [which restricts the maximum intensity of macro-deformation as carbon black-reinforced rubber-like materials violate the incompressibility condition at very large deformations due to debonding of filler particles from the matrix (Gurvich and Fleischman, 2003)], and (ii) junctions in the equivalent network are permanent (with reference to Green and Tobolsky, 1946and Tanaka and Edwards, 1992, this means that the viscoelastic response of the polymer composite is disregarded). Observations depicted in Fig. 5 show that the latter assumption is acceptable from the engineering standpoint for macro-deformations with strain rates of order of and higher (when relaxation of stresses in a test does not exceed 5%).
Acknowledgments
This work was partially supported by The Danish Energy Authority through project ENS–33033–0096.
References
- Baeurle et al., 2005
- A new semi-phenomenological approach to predict the stress relaxation behavior of thermoplastic elastomers
- Polymer, Volume 46, 2005, pp. 4344–4354
- | |
- Bergstrom et al., 2002
- Constitutive modeling of ultra-high molecular weight polyethylene under large-deformation and cyclic loading conditions
- Biomaterials, Volume 23, 2002, pp. 2329–2343
- | |
- Blundell et al., 2002
- Real time SAXS/stress–strain studies of thermoplastic polyurethanes at large strains
- Polymer, Volume 43, 2002, pp. 5197–5207
- | |
- Blundell et al., 2004
- Time-resolved SAXS/stress–strain studies of thermoplastic polyurethanes during mechanical cycling at large strains
- J. Macromol. Sci. Phys., Volume 43 B, 2004, pp. 125–142
- |
- Boyce et al., 2001a
- Deformation of thermoplastic vulcanizates
- J. Mech. Phys. Solids, Volume 49, 2001, pp. 1073–1098
- | |
- Boyce et al., 2001b
- Micromechanisms of deformation and recovery in thermoplastic vulcanizates
- J. Mech. Phys. Solids, Volume 49, 2001, pp. 1323–1342
- Boyce et al., 2001c
- Micromechanics of cyclic softening in thermoplastic vulcanizates
- J. Mech. Phys. Solids, Volume 49, 2001, pp. 1343–1360
- | |
- Datta et al., 1996
- Dynamic mechanical and infra-red spectroscopic studies on interaction between carbon black and ionic thermoplastic elastomer based on maleated EPDM rubber
- Polymer, Volume 37, 1996, pp. 3431–3435
- | |
- Diani et al., 2006a
- A damage directional constitutive model for Mullins effect with permanent set and induced anisotropy
- Eur. J. Mech. A, Volume 25, 2006, pp. 483–496
- | |
- Diani et al., 2006b
- Observation and modeling of the anisotropic visco-hyperelastic behavior of a rubberlike material
- Int. J. Solids Struct., Volume 43, 2006, pp. 3044–3056
- | |
- Drozdov, 2006
- Finite elasticity of thermoplastic elastomers
- Polymer, Volume 47, 2006, pp. 3650–3660
- | |
- Drozdov, 2007
- An unusual elastoplastic response of thermoplastic elastomers at cyclic deformation
- Int. J. Eng. Sci., Volume 45, 2007, pp. 660–678
- | |
- Drozdov and Christiansen, 2006
- Constitutive equations for the nonlinear viscoelastic and viscoplastic behavior of thermoplastic elastomers
- Int. J. Eng. Sci., Volume 44, 2006, pp. 205–226
- | |
- Drozdov and Christiansen, 2007
- Cyclic viscoplasticity of thermoplastic elastomers
- Acta Mech., Volume 194, 2007, pp. 47–65
- |
- Drozdov et al., 2008
- Drozdov, A.D., Jensen, E.A., Christiansen, J.deC., 2008. Thermo-viscoelasticity of polymer melts: experiments and modeling. Acta Mech. 197, 211–245.
- Eceiza et al., 2008
- Thermoplastic polyurethane elastomers based on polycarbonate diols with different soft segment molecular weight and chemical structure: Mechanical and thermal properties
- Polym. Eng. Sci., Volume 48, 2008, pp. 297–306
- |
- Grady and Cooper, 2005
- Thermoplastic elastomers
- The Science and Technology of Rubber, J.E. Mark, B. Erman, F.R. Eirich, 2005, Academic Press, New York, pp. 555–617
- | |
- Green and Tobolsky, 1946
- A new approach to the theory of relaxing polymeric media
- J. Chem. Phys., Volume 14, 1946, pp. 80–92
- |
- Gurvich and Fleischman, 2003
- A simple approach to characterize finite compressibility of elastomers
- Rubber Chem. Technol., Volume 76, 2003, pp. 912–922
- |
- Haupt and Sedlan, 2001
- Viscoplasticity of elastomeric materials: experimental facts and constitutive modelling
- Arch. Appl. Mech., Volume 71, 2001, pp. 89–109
- |
- Holden et al., 2004
- Thermoplastic Elastomers, G. Holden, H.R. Kricheldorf, R.P. Quirk, 2004, Hanser, Munich
- Ihlemann, 2005
- Directional sensitivity of Mullins effect
- Kaut. Gummi Kunstst., Volume 58, 2005, pp. 438–447
- Indukuri and Lesser, 2005
- Comparative deformational characteristics of poly(styrene-b-ethylene-co-butylene-b-styrene) thermoplastic elastomers and cross-linked natural rubber
- Polymer, Volume 46, 2005, pp. 7218–7229
- | |
- Itskov et al., 2006
- Experimental observation of the deformation induced anisotropy of the Mullins effect in rubber
- Kaut. Gummi Kunstst., Volume 59, 2006, pp. 93–96
- Katbab et al., 2000
- Carbon black-reinforced dynamically cured EPDM/PP thermoplastic elastomers. I. Morphology, rheology, and dynamic mechanical properties
- J. Appl. Polym. Sci., Volume 75, 2000, pp. 1127–1137
- |
- Klüppel, 2003
- The role of disorder in filler reinforcement of elastomers on various length scales
- Adv. Polym. Sci., Volume 164, 2003, pp. 1–86
- |
- Kojima et al., 2004
- Morphology–properties relationship of a ternary polymer alloy
- Polym. Eng. Sci., Volume 32, 2004, pp. 1863–1869
- Krijgsman and Gaymans, 2004
- Tensile and elastic properties of thermoplastic elastomers based on PTMO and tetra-amide units
- Polymer, Volume 45, 2004, pp. 437–446
- | |
- Kurian et al., 1995
- Reinforcement of EPDM-based ionic thermoplastic elastomer by carbon black
- Polymer, Volume 36, 1995, pp. 3875–3884
- | |
- Laity et al., 2006
- Morphological behaviour of thermoplastic polyurethanes during repeated deformation
- Macromol. Mater. Eng., Volume 291, 2006, pp. 301–324
- |
- Le et al., 2003
- Time dependent deformation behavior of thermoplastic elastomers
- Polymer, Volume 44, 2003, pp. 4589–4597
- | |
- Long and Sotta, 2006
- Nonlinear and plastic behavior of soft thermoplastic and filled elastomers studied by dissipative particle dynamics
- Macromolecules, Volume 39, 2006, pp. 6282–6297
- |
- Malzbender et al., 2007
- Symmetric shear test of glass–ceramic sealants at SOFC operation temperature
- J. Mater. Sci., Volume 42, 2007, pp. 6297–6301
- |
- Qi and Boyce, 2005
- Stress–strain behavior of thermoplastic polyurethanes
- Mech. Mater., Volume 37, 2005, pp. 817–839
- | |
- Roland et al., 2007
- High strain rate mechanical behavior of polyurea
- Polymer, Volume 48, 2007, pp. 574–578
- | |
- Sarva et al., 2007
- Stress–strain behavior of a polyurea and a polyurethane from low to high strain rates
- Polymer, Volume 48, 2007, pp. 2208–2213
- | |
- Tanaka and Edwards, 1992
- Viscoelastic properties of physically cross-linked networks. Transient network theory
- Macromolecules, Volume 25, 1992, pp. 1516–1523
- |
- Tomita et al., 2007
- Strain-rate-dependent deformation behavior of carbon-black-filled rubber under monotonic and cyclic straining
- Key Eng. Mater., Volume 340–341, 2007, pp. 1017–1024
- |
- Wang and Wang, 2005
- Transient analysis of polymer electrolyte fuel cells
- Electroch. Acta, Volume 50, 2005, pp. 1307–1315
- | |
- Yamauchi et al., 2005
- Structural study of natural rubber thermoplastic elastomers and their composites with carbon black by small-angle neutron scattering and transmission electron microscopy
- Composites A, Volume 36, 2005, pp. 423–429
- | |
- Yi et al., 2006
- Large deformation rate-dependent stress–strain behavior of polyurea and polyurethanes
- Polymer, Volume 47, 2006, pp. 319–329
- | |
- ⁎ Corresponding author. Tel.: +45 72 20 31 42; fax: +45 72 20 31 12.
For further details log on website :
http://www.sciencedirect.com/science/article/pii/S0020768309000419
No comments:
Post a Comment