Skip to main content

Unidirectional Fiber-Reinforced Rubber

  • Chapter
  • First Online:
Book cover Advanced Tire Mechanics

Abstract

The mystery why a tire, just one part of a vehicle, realizes many functions simultaneously can be explained by the fact that a tire has a composite structure made from composite material. Unidirectional fiber-reinforced rubber (UFRR) is the main composite material of a tire. The elastic and viscoelastic properties of UFRR are defined in principal directions, which are the direction of the fiber or cord and the direction perpendicular to the fiber or cord. Applying the rotation matrix to the properties of UFRR in the principal directions, the properties of UFRR can then be calculated in an arbitrary direction. Because UFRR has nonlinear properties with respect to the direction of the applied external force, the angle and width of UFRR used for a tire belt need to be carefully determined in tire design.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 219.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 279.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Problem 1.1.

  2. 2.

    Note 1.1.

  3. 3.

    See Footnote 3.

  4. 4.

    Problem 1.2.

  5. 5.

    Problem 1.3.

  6. 6.

    Note 1.2.

  7. 7.

    See Footnote 5.

  8. 8.

    \(W_{\text{f}} = \frac{{\rho_{\text{f}} V_{\text{f}} }}{{\rho_{\text{m}} V_{\text{m}} + \rho_{\text{f}} V_{\text{f}} }},\;\;\frac{1}{{W_{\text{f}} }} = \frac{{\rho_{\text{m}} V_{\text{m}} }}{{\rho_{\text{f}} V_{\text{f}} }} + 1 = \frac{{\rho_{\text{m}} }}{{\rho_{\text{f}} }}\frac{{1 - V_{\text{f}} }}{{V_{\text{f}} }} + 1.\)

  9. 9.

    Poisson’s ratio for rubber is given by νm = 0.5 owing to incompressibility. This condition yields Gm = Em/{2(1 + νm)} = Em/3.

  10. 10.

    Problem 1.5.

  11. 11.

    Problem 1.6.

  12. 12.

    The relationship [R]−1[T]−1[R] = [T]T is satisfied.

References

  1. T. Hayashi (ed.), Composite Material Engineering (Nikkagiren, Tokyo, 1971) (in Japanese)

    Google Scholar 

  2. S.W. Tsai, Composite Design (Think Composites, Palmetto, 1986)

    Google Scholar 

  3. M. Miki et al., Composite Mechanics (Kyouritsu Shuppan, Tokyo, 1997) (in Japanese)

    Google Scholar 

  4. S.K. Clark (ed.), Mechanics of Pneumatic Tires (U.G. Government Printing Office, Washington, D.C., 1981)

    Google Scholar 

  5. J.D. Walter, Cord-rubber tire composites: theory and application. Rubber Chem. Technol. 51(3), 524–576 (1978)

    Article  Google Scholar 

  6. C.C. Chamis, Simplified composite micromechanics equations for hygral, thermal and mechanical properties. SAMPE Q. 15(3), 14–23 (1984)

    Google Scholar 

  7. R.F. Gibson, Principles of Composite Material Mechanics, 3rd edn. (CRC Press, Boca Raton, 2012)

    Google Scholar 

  8. T. Akasaka, K. Yoshida, Practical Energy Method (Youkendou, 2000) (in Japanese)

    Google Scholar 

  9. B.D. Agarwal, L.J. Broutman, Analysis and Performance of Fiber Composite (Wiley, New York, 1980)

    Google Scholar 

  10. R.M. Jones, Mechanics of Composite Materials, 2nd edn. (Taylor & Francis, Philadelphia, 1999)

    Google Scholar 

  11. S.K. Clark, Theory of the elastic net applied to cord-rubber composites. Rubber Chem. Technol. 56(2), 372–589 (1983)

    Article  Google Scholar 

  12. G. Tangora, Simplified calculations for multi-ply and rubber sheets as a combination of cord-rubber laminating, in Proceedings of International Conference, Moscow (1971)

    Google Scholar 

  13. T. Akasaka, M. Hirano, Approximate elastic constants of fiber reinforced rubber sheet and its composite laminate. Fukugo Zairyo 1(2), 70–76 (1972)

    Google Scholar 

  14. S.K. Clark, Theory of the elastic net applied to cord-rubber composites. Rubber Chem. Technol. 56(2), 372–589 (1983)

    Article  Google Scholar 

  15. K. Kabe, in Study on tire deformation based on structural mechanics. Ph.D. Thesis of Chuo University (1980) (in Japanese)

    Google Scholar 

  16. R. Chandra, et al., Micromechanical damping models for fiber-reinforced composites: a comparative study. Composites Part A 33, 787–796 (2002)

    Google Scholar 

  17. D.A. Saravanos, C.C. Chamis, Unified micromechanics of damping for unidirectional and off-axis fiber composites. J. Compos. Tech. Res. 12, 31–40 (1990)

    Article  Google Scholar 

  18. Y. Nakanishi et al., Estimation method of damping properties for woven fabric composites. Trans. JSME (C) 72(719), 2042–2047 (2006)

    Article  Google Scholar 

  19. Z. Hashin, Analysis of properties of fiber composites with anisotropic heterogeneous materials. J. Appl. Mech. 46, 543–550 (1979)

    Article  Google Scholar 

  20. S.W. Tsai, in Structural behavior of composite materials. NSA-CR-71 (1964)

    Google Scholar 

  21. J.C. Halpin, S.W. Tsai, in Effect of environmental factors on composite materials. AFML-TR-76-423 (1969)

    Google Scholar 

  22. J.D. Eshelby, The determination of the elastic field of an elliptical inclusion and related problems. Proc. R. Soc. Lond. A241, 376–396 (1957)

    MathSciNet  MATH  Google Scholar 

  23. Y.H. Zhao, G.J. Weng, Effective elastic moduli of ribbon-reinforced composites. Trans. Am. Mech. Engr. 57, 158–167 (1990)

    Google Scholar 

  24. R.D. Adams, D.G.C. Bacon, Effect of fibre orientation and laminate geometry on the dynamic properties of CFRP. J. Comp. Mater. 7, 402–428 (1973)

    Article  Google Scholar 

  25. D.X. Lin et al., Prediction and measurement of vibrational damping parameters of carbon and glass fibre-reinforced plastic plates. J. Compos. Mater. 18, 132–152 (1984)

    Article  Google Scholar 

  26. M.R. Maheri, R.D. Adams, Finite element Prediction of modal response of damped layered composite panels. Compos. Sci. and Technol. 55, 13–23 (1995)

    Article  Google Scholar 

  27. K. Terada et al., A method of viscoelastic two scale analyses for FRP. Trans. JSME (A) 75(76), 1674–1687 (2009)

    Article  Google Scholar 

  28. K. Kobayashi et al., Damping vibration analysis of composite materials using mode superposition and homogenization method. J. Jpn. Soc. Compos. Mater. 41(1), 9–18 (2015)

    Article  Google Scholar 

  29. J. Aboudi, Mechanics of Composite Materials. Studies in Applied Mechanics 29 (Elsevier, Amsterdam, 1991)

    Google Scholar 

  30. F. Tabaddor et al., Visocoelastic loss characteristics of cord-rubber composites. Tire Sci. Technol. 14(2), 75–101 (1986)

    Article  Google Scholar 

  31. K. Fujimoto et al., Study on complex modulus of FRR. J. Compos. Mater. 12(4), 163–170 (1986). (in Japanese)

    Google Scholar 

  32. M. Kaliske, H. Rothert, Damping characteristics of unidirectional fibre reinforced polymer composites. Compos. Eng. 5(5), 551–567 (1995)

    Article  Google Scholar 

  33. M. Kaliske, A formulation of elasticity and viscoelasticity for fibre reinforced material at small and finite strains. Comput. Meth. Appl. Mech. Eng. 185, 225–243 (2000)

    Article  Google Scholar 

  34. M. Koishi et al., Homogenization method for analysis of dynamic viscoelastic properties of composite materials. Trans. JSME (A) 62(602), 2270–2275 (1996)

    Article  Google Scholar 

  35. Z. Shida et al., A rolling resistance simulation of tires using static finite element analysis. Tire Sci. Technol. 27(2), 84–105 (1999)

    Article  MathSciNet  Google Scholar 

  36. Y. Ishikawa, Friction of tire—friction on ice and material characteristics. Nippon Gomu Kyokaishi 70(4), 193–203 (1997)

    Article  MathSciNet  Google Scholar 

  37. A. Doi, Trend of recent technologies of tire. Nippon Gomu Kyokaishi 71(9), 588–594 (1998)

    Article  Google Scholar 

  38. T. Hase, K. Yamaguchi, Short fiber reinforcement of TPE. Nippon Gomu Kyokaishi 69(9), 615–623 (1996)

    Article  Google Scholar 

  39. N. Wada, et al., Effect of filled fibers and their orientations on the wear of short fiber reinforced rubber composites. Nippon Gomu Kyokaishi 66(8), 572–584 (1993)

    Google Scholar 

  40. J.C. Halpin, Stiffness and expansion estimates for oriented short fiber composites. J. Compos. Mater. 3, 732–734 (1969)

    Article  Google Scholar 

  41. T. Nishi, Structure and material properties of rubber composites. Nippon Gomu Kyokaishi 57(7), 417–427 (1984)

    Article  Google Scholar 

  42. S. Abrate, The mechanics of short-fiber-reinforced composites: a review. Rubber Chem. Technlol. 59, 384–404 (1986)

    Article  Google Scholar 

  43. M. Ashida, Reinforced fiber (short fiber). Nippon Gomu Kyokaishi 63(11), 694–701 (1990)

    Article  Google Scholar 

  44. M. Ashida et al., Effect of fiber length on the mechanical and dynamical properties of PET-CR composite. Nippon Gomu Kyokaishi 60(3), 158–164 (1987)

    Article  Google Scholar 

  45. N. Yoshida et al., Effect of damping on earthquake response of ground and its accuracy. J. Jpn. Assoc. Earthquake Eng. 6(4), 55–73 (2006)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yukio Nakajima .

Appendices

Appendix: Viscoelasticity

Using Boltzmann’s superposition principle, the stress response σ(t) with respect to the loading history of the strain ε(t) is expressed by

$$\sigma (t) = \int\limits_{ - \infty }^{t} {E(t - \hat{\tau })\frac{{{\text{d}}\varepsilon (\hat{\tau })}}{{{\text{d}}\hat{\tau }}}} {\text{d}}\hat{\tau }.$$
(1.148)

The nondimensional relaxation function ϕ(t) is introduced as

$$\phi (t) = E(t)/E_{\infty } - 1,$$
(1.149)

where E is the elastic modulus at t = ∞. The substitution of Eq. (1.148) into Eq. (1.149) yields

$$\sigma (t) = \sigma_{\infty } (t) + \int\limits_{ - \infty }^{t} {\phi (t - \hat{\tau })\frac{{{\text{d}}\sigma_{\infty } (\hat{\tau })}}{{{\text{d}}\hat{\tau }}}} {\text{d}}\hat{\tau } = \sigma_{\infty } (t) - \int\limits_{0}^{\infty } {\phi (\tau )\frac{{{\text{d}}\sigma_{\infty } (t - \tau )}}{{{\text{d}}\tau }}} {\text{d}}\tau ,$$
(1.150)

where σ = Eε(t) and \({\tau } = {t} - \hat{\tau }.\)

Suppose that the strain ε harmonically oscillates with the angular velocity ω and ε is expressed by

$$\varepsilon (t) = \varepsilon^{*} {\text{e}}^{{{\text{i}}\omega t}} .$$
(1.151)

The substitution of Eq. (1.151) into Eq. (1.150) yields

$$\sigma (t) = E_{\infty } \varepsilon (t) - \int\limits_{0}^{\infty } {\phi (\tau )\frac{\text{d}}{{{\text{d}}\tau }}} \left( {E_{\infty } \varepsilon^{*} {\text{e}}^{{{\text{i}}\omega \left( {t - \tau } \right)}} } \right){\text{d}}\tau = E_{\infty } \left( {1 + {\text{i}}\omega \int\limits_{0}^{\infty } {\phi (\tau ){\text{e}}^{{ - {\text{i}}\omega \tau }} } {\text{d}}\tau } \right)\varepsilon (t).$$
(1.152)

Considering that ϕ(τ) = 0 is satisfied when τ < 0, the Fourier transformation of ϕ(t) is expressed as

$$\phi^{*} (\omega ) = \int\limits_{0}^{\infty } {\phi (\tau ){\text{e}}^{{ - {\text{i}}\omega \tau }} } {\text{d}}\tau .$$
(1.153)

Equation (1.152) is rewritten as

$$\sigma (t) = E^{*} (\omega )\varepsilon (t),$$
(1.154)

where \(E^{*} \left( \omega \right)\) is given by

$$E^{*} (\omega ) = E_{\infty } \left( {1 + {\text{i}}\omega \phi^{*} (\omega )} \right).$$
(1.155)

Expressing ϕ*(ω) in complex form, Eq. (1.155) can be rewritten as

$$E^{*} (\omega ) = E_{\infty } \left( {1 - \omega \phi^{\prime \prime } } \right) + {\text{i}}E_{\infty } \omega \phi^{{\prime }} = E^{\prime} + {\text{i}}E^{\prime\prime},$$
(1.156)

where \(\phi^{\prime}\) and \(\phi^{\prime\prime}\) are, respectively, real and imaginary parts of ϕ*(ω). \(E^{\prime}\) and \(E^{\prime\prime}\) are, respectively, referred to as the storage modulus and loss modulus. The loss tangent tan δ is defined by

$$\tan \delta = E^{\prime\prime}/E^{\prime}.$$
(1.157)

Identification of the Damping Matrix [45]

The vibration problem of linear elastic materials can be solved by following one of two approaches of the finite element method. One is the direct approach, which is an exact method, but the computational cost is high and a method of identifying the damping matrix has not been established. The other approach is the modal superposition method, which is a kind of approximation method that has low computational cost. The damping coefficient can be determined from experimental modal analysis. Yoshida [45] proposed identifying the damping matrix for the direct method using modal damping coefficients.

The equation of motion for the structure with damping is given by

$$\left[ M \right]\left\{ {\ddot{u}} \right\} + \left[ C \right]\left\{ {\dot{u}} \right\} + \left[ K \right]\left\{ u \right\} = \left\{ F \right\},$$
(1.158)

where {u} is the displacement vector, the dot indicates the time derivative, and [M], [C], [K] and {F} are the mass matrix, damping matrix, stiffness matrix and external force. When the damping and external force are zero in Eq. (1.158), the structure freely vibrates. The eigenvalue problem is thus defined, where the r-th real eigenvector is expressed by {Xr}. The solution {u} to Eq. (1.158) can be expressed as the superposition of orthogonal vectors {Xr}:

$$\left\{ u \right\} = \left[ {\left\{ {X_{1} } \right\}\left\{ {X_{2} } \right\} \ldots \left\{ {X_{r} } \right\} \ldots \left\{ {X_{n} } \right\}} \right]\left\{ q \right\} = \left[ \xi \right]\left\{ q \right\},$$
(1.159)

where {q} is the modal displacement, [ξ] is the matrix for which the eigenvector {Xr} is in the column. Substituting Eq. (1.159) into Eq. (1.158) and multiplying by [ξ]T from the left yields

$$\begin{aligned} & \left[ m \right]\left\{ {\ddot{u}} \right\} + \left[ c \right]\left\{ {\dot{u}} \right\} + \left[ k \right]\left\{ u \right\} = \left[ \xi \right]^{\text{T}} \left\{ F \right\} \\ & \left[ m \right] = \left[ \xi \right]^{\text{T}} \left[ M \right]\left[ \xi \right] \\ & \left[ c \right] = \left[ \xi \right]^{\text{T}} \left[ C \right]\left[ \xi \right] \\ & \left[ k \right] = \left[ \xi \right]^{\text{T}} \left[ K \right]\left[ \xi \right], \\ \end{aligned}$$
(1.160)

where mass matrix [m] and stiffness matrix [k] are diagonal matrices because of the orthogonality of eigenvectors. The r-th diagonal element mr and kr are referred to as the modal mass and modal stiffness and are expressed as

$$\begin{aligned} & \left[ m \right]\left\{ {\ddot{u}} \right\} + \left[ c \right]\left\{ {\dot{u}} \right\} + \left[ k \right]\left\{ u \right\} = \left[ \xi \right]^{\text{T}} \left\{ F \right\} \\ & m_{r} = \left\{ {X_{r} } \right\}^{\text{T}} \left[ M \right]\left\{ {X_{r} } \right\} \\ & k_{r} = \left\{ {X_{r} } \right\}^{\text{T}} \left[ K \right]\left\{ {X_{r} } \right\}. \\ \end{aligned}$$
(1.161)

Meanwhile, the eigenvector {Xr} does not satisfy orthogonality with the damping matrix [C], and the third equation [c] of Eq. (1.160) is not a diagonal matrix in general. If [C] is expressed by the linear combination of [K] and [M], it is referred to as Rayleigh damping and the matrix [c] becomes a diagonal matrix. However, because Rayleigh damping has a damping matrix with two parameters, the properties of the damping matrix cannot be expressed satisfactory. Yoshida assumed that [c] is a diagonal matrix in which the diagonal element is the modal damping cr and then expressed [C] by \(c_{r}\). This assumption gives a relationship between cr and [C]:

$$c_{r} = \left\{ {X_{r} } \right\}^{\text{T}} \left[ C \right]\left\{ {X_{r} } \right\}.$$
(1.162)

Here, cr is expressed by

$$c_{r} = 2\omega_{r} \zeta_{r} m_{r} ,$$
(1.163)

where ζr is the modal damping ratio.

The substitution of Eq. (1.163) into Eq. (1.162) yields

$$\left[ {2\omega_{r} \zeta_{r} m_{r} } \right] = \left[ \xi \right]^{\text{T}} \left[ C \right]\left[ \xi \right],$$
(1.164)

where [2ωrζrmr] is a diagonal matrix in which the r-th diagonal element is 2ωrζrmr. Using the second equation of Eq. (1.160), Eq. (1.164) can be transformed to

$$\left[ C \right] = \left[ M \right]\left[ \xi \right]\left[ {2\omega_{r} \zeta_{r} /m_{r} } \right]\left[ \xi \right]^{\text{T}} \left[ M \right],$$
(1.165)

where [2ωrζr/mr] is a diagonal matrix in which the r-th diagonal element is 2ωrζr/mr . Note that [ξ]−1 cannot be calculated because [ξ] is not a square matrix.

Physical Meaning of Modal Damping [28]

When [c] is a diagonal matrix, the equations given as Eq. (1.161) can be rewritten as n one-dimensional equations of motion:

$$m_{r} \ddot{q}_{r} + 2\omega_{r} \zeta_{r} m_{r} \dot{q}_{r} + k_{r} q_{r} = f_{r} ,$$
(1.166)

where the undamped natural frequency ωr and generalized external force fr are

$$\begin{aligned} \omega_{r} & = \sqrt {k_{r} /m_{r} } \\ f_{r} & = \left\{ {X_{r} } \right\}^{\text{T}} \left\{ F \right\}. \\ \end{aligned}$$
(1.167)

When fr is zero in Eq. (1.166) and the condition ζr < 1 is satisfied, the solution is damped vibration. Suppose that the period is T. Multiplying both sides of Eq. (1.166) by dqr and integrating over one period of time, we obtain

$$\begin{aligned} & \int\limits_{0}^{T} {\frac{{{\text{d}}K_{\text{en}} }}{{{\text{d}}t}}{\text{d}}t} + \int\limits_{0}^{T} {\frac{{{\text{d}}U}}{{{\text{d}}t}}{\text{d}}t} = - 2\omega_{r} \zeta_{r} m_{r} \int\limits_{0}^{T} {\dot{q}_{r}^{2} {\text{d}}t} \\ & K_{\text{en}} = \frac{1}{2}m_{r} \dot{q}_{r}^{2} \\ & U = \frac{1}{2}k_{r} q_{r}^{2} , \\ \end{aligned}$$
(1.168)

where Ken and U are, respectively, the kinetic energy and strain energy. Equation (1.168) gives the energy conservation in a period. Assuming the initial displacement is qr0 at t = 0, the solution to (1.166) is given by

$$\begin{aligned} q_{r} & = q_{r0} {\text{e}}^{{ - \omega_{r} \zeta t}} \cos \omega_{rd} t \\ \omega_{{r{\text{d}}}} & = \omega_{r} \sqrt {1 - \zeta_{r}^{2} } , \\ \end{aligned}$$
(1.169)

where ωrd is the natural frequency of the damped vibration of the r-th mode. Substituting Eq. (1.169) into the right side of Eq. (1.168), applying a Taylor series expansion with respect to ζr and then neglecting more than the quadratic term of ζr, we obtain

$$2\omega_{r} \zeta_{r} m_{r} \int\limits_{0}^{T} {\dot{q}_{r}^{2} {\text{d}}t} \approx 4\pi \zeta_{r} \left( {\frac{1}{2}k_{r} q_{r0}^{2} } \right) = 4\pi \zeta_{r} E_{0} ,$$
(1.170)

where E0 is the initial strain energy. Because the first term on the left side of Eq. (1.168) is the energy −ΔE to lose in a period, from Eq. (1.168) and (1.170), we obtain

$$\Delta E = 4\pi \zeta_{r} E_{0} .$$
(1.171)

Equation (1.171) shows that the dissipation energy in a period, ΔE, is proportional to the initial strain energy E0 and the modal damping ratio ζr is equivalent to the ratio of the two energies. When the material is linear viscoelastic, the constitutive equation is Eq. (1.154). When the strain is assumed to be harmonic vibration expressed by Eq. (1.151), and the dissipation energy Δe in a period per unit volume is the same as the work of internal stress, this is expressed by

$$\Delta e = \oint {\text{Re} \left( {\left\{ {\sigma_{r} } \right\}} \right)^{\text{T}} \text{Re} \left( {\left\{ {{\text{d}}\varepsilon_{r} } \right\}} \right)} = \oint {\text{Re} \left( {\left[ {E^{*} } \right]\left\{ {\sigma_{r} } \right\}} \right)^{\text{T}} \cdot \text{Re} \left( {\left\{ {\dot{\varepsilon }_{r} } \right\}} \right)} {\text{d}}t = \pi \left\{ {\varepsilon_{r} } \right\}^{\text{T}} \left[ {E^{\prime\prime}} \right]\left\{ {\varepsilon_{r} } \right\},$$
(1.172)

where Re indicates the real part of the complex number. Integrating Δe in the volume of the structure, the energy loss in a period ΔE and the initial strain energy E0 are expressed by

$$\begin{aligned} \Delta E & = \pi \int\limits_{V} {\left\{ {\varepsilon_{r} } \right\}^{\text{T}} \left[ {E^{\prime\prime}} \right]\left\{ {\varepsilon_{r} } \right\}{\text{d}}V} \\ E_{0} & = \frac{1}{2}\int\limits_{V} {\left\{ {\varepsilon_{r} } \right\}^{\text{T}} \left[ {E^{\prime}} \right]\left\{ {\varepsilon_{r} } \right\}{\text{d}}V} . \\ \end{aligned}$$
(1.173)

Equations (1.171) and (1.173) give the modal damping ratio ζr expressed by

$$\zeta_{r} = \frac{1}{2}\frac{{\int_{V} {\left\{ {\varepsilon_{r} } \right\}^{\text{T}} \left[ {E^{\prime\prime}} \right]\left\{ {\varepsilon_{r} } \right\}{\text{d}}V} }}{{\int_{V} {\left\{ {\varepsilon_{r} } \right\}^{\text{T}} \left[ {E^{\prime}} \right]\left\{ {\varepsilon_{r} } \right\}{\text{d}}V} }}.$$
(1.174)

Hence, if the complex elastic matrix \([E^{*} ]\) is known, ζr can be calculated from the strain energy distribution of the mode corresponding to the real eigenvector.

Notes

Note 1.1 Incompressibility of Rubber

Rubber has specific properties that the volume is maintained under external forces. Referring to Fig. 1.35a, the incompressibility is expressed by

$$\left( {1 + \varepsilon_{x} } \right)l_{x} \left( {1 + \varepsilon_{y} } \right)l_{y} \left( {1 + \varepsilon_{z} } \right)l_{z} = l_{x} l_{y} l_{z} .$$

The low-strain assumptions εx, εy, εz ≪ 1 yield

$$\varepsilon_{x} + \varepsilon_{y} + \varepsilon_{z} = 0.$$

Figure 1.35b shows the deformation of the element where the external force is applied in the z-direction. When the strain along the z-axis is ε, strains along the x- and y-axes can be expressed by ε and Poisson’s ratio ν. Referring to Fig. 1.35b, the incompressibility can be expressed by

$$\left( {1 - \nu \varepsilon } \right)l_{x} \left( {1 - \nu \varepsilon } \right)l_{y} \left( {1 + \varepsilon } \right)l_{z} = l_{x} l_{y} l_{z} .$$

The assumption of low strain ε ≪ 1 yields

$$\left( {1 - \nu \varepsilon } \right)^{2} \left( {1 + \varepsilon } \right) \cong \left( {1 - 2\nu \varepsilon } \right)\left( {1 + \varepsilon } \right) \cong 1 + \left( {1 - 2\nu } \right)\varepsilon = 1.$$

Poisson’s ratio for incompressible materials is ν = 1/2 from the above equation.

Fig. 1.35
figure 35

Incompressibility of rubber

Note 1.2

When −1 < cos 2θ < 1 is satisfied, the extreme value is located away from θ = 0° and θ = 90°. Because \(\frac{{E_{T} - E_{L} }}{{E_{L} E_{T} /G_{LT} - \left( {E_{L} + E_{T} + 2\nu_{L} E_{T} } \right)}} < 1\) is satisfied, GLT > EL/{2(1 + νL)} is obtained considering the relation EL − ET > 0.

In the case of FRR, from Eq. (1.98), the elastic properties can be expressed as

$$\begin{aligned} & G_{LT} = \frac{{E_{\text{m}} }}{3} \\ & \frac{{E_{L} }}{{2(1 + \nu_{L} )}} = \frac{{E_{\text{f}} V_{\text{f}} }}{3} \\ & \frac{{E_{T} }}{{2(1 + \nu_{T} )}} = \frac{2}{3}E_{\text{m}} . \\ \end{aligned}$$

Because the condition GLT < ET/{2(1 + νT)} is satisfied for FRR, Ex for FRR has an extreme value within 0 < θ < 90°.

Rights and permissions

Reprints and permissions

Copyright information

© 2019 Springer Nature Singapore Pte Ltd.

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Nakajima, Y. (2019). Unidirectional Fiber-Reinforced Rubber. In: Advanced Tire Mechanics. Springer, Singapore. https://doi.org/10.1007/978-981-13-5799-2_1

Download citation

  • DOI: https://doi.org/10.1007/978-981-13-5799-2_1

  • Published:

  • Publisher Name: Springer, Singapore

  • Print ISBN: 978-981-13-5798-5

  • Online ISBN: 978-981-13-5799-2

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics