Fact-checked by Grok 2 weeks ago

Transverse isotropy

Transverse isotropy refers to a specific form of material in which properties are identical in all directions within a plane of , known as the , but exhibit distinct behavior along the axis perpendicular to that plane, termed the axis of . This reduces the complexity of describing the material's response compared to general , making it a common idealization for layered or fibrous structures in . In the context of , transversely isotropic materials are characterized by five independent elastic constants, as opposed to the two for fully isotropic materials or nine for general orthotropic materials. These constants typically include moduli and Poisson's ratios for both the and the axial direction, along with a coupling the two. The for such materials contains 12 nonzero terms but reflects the rotational invariance around the symmetry axis, simplifying stress-strain relations under . Transversely isotropic models are widely applied in to represent unidirectional fiber-reinforced composites, such as carbon fiber-epoxy systems, where fibers align parallel to the axis. Other examples include certain piezoelectric ceramics like PZT-4, biological tissues such as , and geological formations like , which display this due to their microstructure. Experimental often relies on ultrasonic wave propagation or to determine these constants, enabling accurate predictions of deformation and failure in engineering designs.

Fundamentals

Definition

Transverse isotropy refers to a type of in which physical properties are identical in every direction within a specific , known as the , but may vary along the direction perpendicular to that , termed the axis. This configuration results in effective hexagonal and represents an intermediate state between full and more complex forms of . The concept of transverse isotropy originated in 19th-century studies of anisotropic media within crystal physics, building on foundational work in elasticity by researchers such as , who contributed to the understanding of wave propagation in anisotropic solids. Further developments in the mathematical description of crystal symmetries were advanced by in his investigations of elastic properties in crystals. The specific term "transversely isotropic" gained prominence in the early through A. E. H. Love's comprehensive on elasticity. In contrast to fully isotropic materials, which exhibit uniform properties in and require only two independent elastic constants, transversely isotropic materials possess a single preferred that distinguishes properties along and to it. They differ from orthotropic materials, which feature three mutually planes of symmetry and thus nine independent elastic constants. For , transverse isotropy simplifies the general anisotropic case—from 21 independent constants—to five, reflecting the reduced symmetry constraints.

Symmetry Characteristics

Transverse isotropy is defined by a group consisting of all rotations about a distinguished , conventionally taken as the z-, combined with reflections across any containing that . This group, often denoted as the O(2) acting in the , ensures that material properties are invariant under arbitrary rotations within the x-y and under reflections through vertical planes. Such symmetries mirror those of a , where the represents the unique direction of , and the perpendicular exhibits full . The invariance properties extend to 180-degree rotations about any axis lying in the x-y plane, which can be derived from compositions of the primary rotations and reflections, preserving the overall material response. This geometric arrangement distinguishes the symmetry axis from the isotropic plane, where directional properties are identical in all orientations. In , transverse isotropy corresponds directly to the , characterized by a six-fold rotational axis (approximated as continuous for macroscopic descriptions) along the c-axis, leading to equivalent physical behaviors in layered or fibrous structures. Visually, the can be conceptualized as a cylindrical form, with the z-axis serving as the infinite-fold rotation axis and the surrounding x-y as the locus of , unbounded by discrete facets. This framework underpins the reduction of the general 21 constants to five independent parameters in , highlighting the efficiency of the in constraining material behavior.

Material Examples

Natural and Geological Materials

Transverse is commonly observed in sedimentary rocks such as shales, where fine layering from depositional processes creates a of parallel to the and an of perpendicular to it. This symmetry arises because deposits materials in layers, resulting in isotropic within the while exhibiting distinct vertical due to varying alignment and compaction. For instance, shales often display vertical transverse (VTI), with that differ significantly along the vertical compared to the directions. In fractured reservoirs, transverse isotropy emerges from the alignment of cracks or s, which preferentially orient due to tectonic stresses or flow, creating effective symmetry in the plane perpendicular to the dominant fracture direction. This leads to horizontal transverse isotropy (HTI) when fractures are vertical and aligned, influencing permeability and mechanical stability in hydrocarbon-bearing formations. The background rock fabric, combined with fracture , further modulates this , making it a key factor in . Among natural materials, exhibits transverse isotropy due to the aligned fibers along the longitudinal growth direction, with radial and tangential planes showing near- from annual growth rings and vascular patterns. This results from biological growth mechanisms in trees, where cells elongate preferentially along the to support vertical load-bearing, while transverse directions remain more uniform due to circumferential expansion. Similarly, , a formed under high-pressure metamorphic conditions, displays transverse isotropy from its stacked hexagonal layers, which bond weakly between planes, yielding in the basal plane and strong perpendicular to it. Observational evidence for transverse isotropy in the comes from seismic anisotropy studies, where VTI media are inferred from variations in P- and S-wave velocities, often attributed to aligned microstructures in sedimentary layers or aligned minerals in the middle and lower crust. Such is detected through shear-wave splitting and azimuthal variations in crustal seismic profiles, providing insights into depositional history and tectonic alignment without direct sampling.

Engineered and Biological Materials

Unidirectional fiber-reinforced composites, such as those with embedded in a , exhibit transverse isotropy due to the aligned orientation of s along a single direction, resulting in isotropic properties in the plane perpendicular to the fiber axis. This symmetry arises from the transverse isotropy of the individual s combined with the , where the fiber volume fraction and alignment are intentionally engineered to enhance longitudinal stiffness and strength for applications like structures. Piezoelectric ceramics, including (PZT-4) and , display transverse isotropy following poling, a process that applies an to align ferroelectric domains along a preferred , rendering properties isotropic in the perpendicular plane. This engineered optimizes electromechanical coupling for sensors and actuators, with the poling direction serving as the of symmetry. In biological materials, cortical exhibits transverse isotropy with the symmetry axis aligned along the longitudinal direction, arising from the oriented microstructure of osteons (Haversian systems). This alignment provides enhanced stiffness and strength along the bone's length while maintaining in the transverse plane. Soft tissues like arterial walls demonstrate transverse isotropy from the preferential alignment and dispersion of collagen fibers in the tangential plane, providing enhanced stiffness against circumferential and axial loads while maintaining isotropy out-of-plane. This physiological , evolved for efficient containment and pulsatile flow resistance, features symmetric fiber families (e.g., at ±48° to the circumferential direction in human ) with controlled in-plane dispersion. Similarly, tendons exhibit transverse isotropy through helical collagen fiber arrangements around tenocytes, enabling high tensile stiffness along the load-bearing direction and coupled compressive responses for force transmission from muscle to . Verification of transverse isotropy in biological soft tissues, such as , often employs shear wave elastography, which measures directional differences in shear wave speeds (e.g., higher along fibers at ~26 kPa elasticity versus ~8 kPa across) to confirm anisotropic viscoelastic properties.

Mathematical Framework

Symmetry Transformation Matrix

Transverse isotropy is characterized by a that includes all s about a distinguished , typically taken as the z-axis, and reflections through planes containing this axis. These transformations leave the material properties invariant, ensuring that physical responses are identical in the perpendicular to the axis. The rotation matrices corresponding to arbitrary rotations by an angle θ about the z-axis take the form R(\theta) = \begin{pmatrix} \cos\theta & -\sin\theta & 0 \\ \sin\theta & \cos\theta & 0 \\ 0 & 0 & 1 \end{pmatrix}, which generates the special orthogonal group SO(2) in the xy-plane. Reflection matrices, which account for the full orthogonal group O(2), include, for example, reflection across the xz-plane: S_x = \begin{pmatrix} 1 & 0 & 0 \\ 0 & -1 & 0 \\ 0 & 0 & 1 \end{pmatrix}, with similar forms for reflections across other vertical planes obtained by composing with rotations. For a material property tensor T of second rank to exhibit transverse isotropy, it must satisfy the invariance T' = R T R^T = T under any such R, and analogously under reflections, where T' is the transformed tensor. This extends to higher-rank tensors, such as the fourth-rank , via the appropriate tensor rules, ensuring the overall material response remains unchanged. The invariance under the continuous set of rotations implies that, in coordinates aligned with the symmetry axis, the tensor components must be independent of the azimuthal angle in the , leading to a diagonal or block-diagonal form where the transverse directions are indistinguishable. Specifically, applying the for arbitrary θ enforces equality of the in-plane components (e.g., T_{xx} = T_{yy} and T_{xy} = 0) and zeros certain off-diagonal terms, deriving the standard transversely isotropic form without explicit enumeration of all elements. This symmetry class positions transverse isotropy intermediate among elastic anisotropy classes: it possesses more independent constants than , which requires only two (e.g., λ and μ), but fewer than orthotropy, which has nine. For the , transverse isotropy thus demands five independent constants, reflecting the reduced compared to full isotropy while avoiding the three orthogonal planes of orthotropy.

Elasticity Tensor

In transversely isotropic materials, the fourth-rank C_{ijkl} exhibits specific symmetries arising from the plane of isotropy, typically taken as the x_1-x_2 plane with the axis of symmetry along x_3. The tensor components satisfy C_{ijkl} = C_{jikl} = C_{ijlk} = C_{klij} due to the inherent symmetries of and , and the transverse isotropy imposes additional equalities such as C_{1111} = C_{2222}, C_{1122} = C_{2211}, C_{1133} = C_{2233}, C_{3311} = C_{3322}, C_{1313} = C_{2323}, C_{1212} = C_{1221}, and all other components are zero. To facilitate computations, the elasticity tensor is often represented in Voigt notation as a symmetric 6×6 stiffness matrix \mathbf{C}, relating the six stress components \sigma_i to the six strain components \varepsilon_j via \boldsymbol{\sigma} = \mathbf{C} \boldsymbol{\varepsilon}. For transverse isotropy, this matrix has only five independent components: C_{11} = C_{22}, C_{12}, C_{13} = C_{23}, C_{33}, and C_{44} = C_{55}, with the sixth derived as C_{66} = (C_{11} - C_{12})/2. The explicit form is \mathbf{C} = \begin{pmatrix} C_{11} & C_{12} & C_{13} & 0 & 0 & 0 \\ C_{12} & C_{11} & C_{13} & 0 & 0 & 0 \\ C_{13} & C_{13} & C_{33} & 0 & 0 & 0 \\ 0 & 0 & 0 & C_{44} & 0 & 0 \\ 0 & 0 & 0 & 0 & C_{44} & 0 \\ 0 & 0 & 0 & 0 & 0 & C_{66} \end{pmatrix}. The compliance tensor \mathbf{S}, the inverse of \mathbf{C}, relates strains to stresses via \boldsymbol{\varepsilon} = \mathbf{S} \boldsymbol{\sigma} and shares the same symmetry structure, with five independent components: S_{11} = S_{22}, S_{12}, S_{13} = S_{23}, S_{33}, and S_{44} = S_{55}, where S_{66} = 2(S_{11} - S_{12}). Its matrix form is \mathbf{S} = \begin{pmatrix} S_{11} & S_{12} & S_{13} & 0 & 0 & 0 \\ S_{12} & S_{11} & S_{13} & 0 & 0 & 0 \\ S_{13} & S_{13} & S_{33} & 0 & 0 & 0 \\ 0 & 0 & 0 & S_{44} & 0 & 0 \\ 0 & 0 & 0 & 0 & S_{44} & 0 \\ 0 & 0 & 0 & 0 & 0 & S_{66} \end{pmatrix}. The explicit relations between \mathbf{S} and \mathbf{C} components are obtained by matrix inversion, yielding expressions such as S_{11} = \frac{ C_{11} C_{33} - C_{13}^2 }{ (C_{11} - C_{12}) [ (C_{11} + C_{12}) C_{33} - C_{13}^2 ] }. These symmetries reduce the general anisotropic elasticity tensor, which has 21 independent constants due to its major and minor symmetries, to just five for transverse isotropy by imposing equalities across the components invariant under rotations about the symmetry axis. This elimination of 16 constants simplifies analysis while capturing the essential directional dependence.

Linear Elasticity

Conditions for Symmetry

Transverse isotropy in materials requires specific microstructural arrangements that establish a plane of isotropy perpendicular to a unique symmetry axis. This symmetry emerges when structural elements, such as fibers, layers, or inclusions, are aligned along the preferred axis while exhibiting statistical uniformity and random orientation within the transverse plane. For instance, in unidirectional fiber-reinforced composites, the fibers are oriented parallel to the axis, with their cross-sectional distribution homogenized across the plane to ensure isotropic behavior in all directions normal to the axis. Theoretically, a exhibits transverse isotropy if its constitutive response remains invariant under the operations of its , which includes all rotations about the preferred and reflections through any containing that . This group is a of the Orth(3), and the material's energy function or stress-strain relation must satisfy invariance conditions, such as W(Q C Q^T) = W(C), where Q belongs to the and C is the right Cauchy-Green deformation tensor. Any deviation from this invariance, such as preferred orientations in the or additional mirror planes, reduces the to lower classes like monoclinic or orthorhombic. Experimentally, transverse isotropy is verified by confirming the presence of exactly five independent elastic constants, which distinguish it from full isotropy (two constants) or higher anisotropy. Ultrasonic testing measures shear and compressional wave velocities along the symmetry axis, perpendicular to it, and at intermediate angles (e.g., 45°), allowing computation of these constants from travel times and sample dimensions under controlled stress conditions. Similarly, X-ray diffraction assesses microstructural alignment by analyzing preferred orientations of crystallites or minerals, such as clay platelets parallel to bedding planes, which corroborate the transverse symmetry through pole figure distributions. These methods link the observed elastic tensor form—characterized by five nonzero components—to the underlying symmetry. These conditions presuppose a linear regime, where small deformations preserve the ; in nonlinear responses, such as those in hyper materials under finite strains, the initial transverse isotropy may evolve or break due to microstructural rearrangements or path-dependent hardening.

Constitutive Relations and Engineering Moduli

In , the constitutive relations for transversely isotropic (TI) materials are expressed through in , relating the vector \boldsymbol{\sigma} to the vector \boldsymbol{\epsilon} via the [\mathbf{C}]: \boldsymbol{\sigma} = [\mathbf{C}] \boldsymbol{\epsilon}. For TI with the axis of isotropy along the 3-direction, the 6×6 has five independent components and takes the form [\mathbf{C}] = \begin{pmatrix} C_{11} & C_{12} & C_{13} & 0 & 0 & 0 \\ C_{12} & C_{11} & C_{13} & 0 & 0 & 0 \\ C_{13} & C_{13} & C_{33} & 0 & 0 & 0 \\ 0 & 0 & 0 & C_{44} & 0 & 0 \\ 0 & 0 & 0 & 0 & C_{44} & 0 \\ 0 & 0 & 0 & 0 & 0 & C_{66} \end{pmatrix}, where C_{66} = (C_{11} - C_{12})/2. The engineering moduli are derived from the inverse compliance matrix [\mathbf{S}] = [\mathbf{C}]^{-1}, which relates strain to stress: \boldsymbol{\epsilon} = [\mathbf{S}] \boldsymbol{\sigma}. The longitudinal Young's modulus E_L (along the symmetry axis) is E_L = 1/S_{33}, the transverse Young's modulus E_T (in the isotropy plane) is E_T = 1/S_{11}, the in-plane shear modulus G_T is G_T = 1/S_{66}, and the axial shear modulus G_L is G_L = C_{44}. The Poisson's ratios include the transverse-transverse ratio \nu_{TT} = -S_{12}/S_{11} and the longitudinal-transverse ratio \nu_{LT} = -S_{13}/S_{33}, with the reciprocal \nu_{TL} = \nu_{LT} (E_T / E_L) enforced by symmetry. The stiffness constants C_{ij} are linked to these engineering moduli through explicit relations obtained by inverting the compliance matrix. For instance, C_{11} = E_T (1 - \nu_{TT}^2) / \Delta, C_{12} = E_T (\nu_{TT} + \nu_{LT} \nu_{TL}) / \Delta, C_{13} = E_L \nu_{LT} / \Delta, and C_{33} = E_L (1 - \nu_{TT}^2) / \Delta, where \Delta = 1 - \nu_{TT}^2 - 2 \nu_{LT} \nu_{TL} (1 + \nu_{TT}) is the determinant factor ensuring positive definiteness. Additionally, C_{44} = G_L and C_{66} = G_T. These expressions facilitate practical computation of anisotropic responses in TI materials. The strain energy density function for TI materials under is the U = \frac{1}{2} C_{ijkl} \epsilon_{ij} \epsilon_{kl}, which simplifies due to the reduced number of independent C_{ijkl} components in the TI . This form ensures the material's and reflects the uncoupled energy modes aligned with the , such as volumetric, deviatoric, and contributions in the .

Applications in Geophysics

Backus Upscaling (Long-Wavelength Approximation)

The Backus upscaling method, also known as Backus averaging, provides a theoretical for approximating a stack of thin, horizontally layered anisotropic media as an effective homogeneous vertically transversely isotropic (VTI) medium in the long- seismic , where the significantly exceeds the individual layer thicknesses. This approach is particularly relevant in for modeling propagation through finely layered formations, such as sedimentary basins or the , by deriving effective elastic properties that capture the overall anisotropic response without resolving fine-scale heterogeneities. Developed by George E. Backus in specifically for interpreting elastic anisotropy in the Earth's interior, the method assumes a periodic or statistically stationary sequence of horizontal layers with transversely isotropic or isotropic properties varying primarily in the vertical direction. Key assumptions include fine-scale layering where layer thicknesses are much smaller than the seismic wavelength (ensuring the long-wavelength approximation holds), perfect bonding between layers with no slip (implying continuity of displacements and tractions), and a quasi-static regime neglecting dynamic inertial effects over the averaging scale. These conditions ensure that the effective medium behaves as if the layers are in a state of uniform or appropriate to the wave mode, leading to a VTI with the aligned perpendicular to the layering. The effective stiffness constants in the Voigt notation for the VTI medium are obtained through weighted averages over the layer thicknesses, where the angle brackets \langle \cdot \rangle denote thickness-weighted averages (e.g., for discrete layers, \langle f \rangle = \sum h_k f_k with h_k the fractional thickness of layer k). The vertical normal stiffness \tilde{C}_{33} is the harmonic average reflecting series stacking for vertical compression: \tilde{C}_{33} = \left\langle \frac{1}{C_{33}} \right\rangle^{-1} Similarly, the shear stiffness for vertically polarized shear waves \tilde{C}_{44} follows a harmonic average due to uniform shear stress across layers: \tilde{C}_{44} = \left\langle \frac{1}{C_{44}} \right\rangle^{-1} The in-plane shear stiffness \tilde{C}_{66} is the arithmetic average, as layers act in parallel for horizontally polarized shear waves: \tilde{C}_{66} = \langle C_{66} \rangle The cross-coupling constant \tilde{C}_{13} accounts for the interaction between vertical and horizontal normal stresses: \tilde{C}_{13} = \tilde{C}_{33} \left\langle \frac{C_{13}}{C_{33}} \right\rangle Finally, the horizontal normal stiffness \tilde{C}_{11} incorporates both the averaged in-plane response and a correction from the coupling term: \tilde{C}_{11} = \left\langle C_{11} - \frac{C_{13}^2}{C_{33}} \right\rangle + \tilde{C}_{33}^2 \left\langle \frac{C_{13}}{C_{33}} \right\rangle^2 These expressions are derived by enforcing equality between the average stresses and strains in the layered stack and the effective medium under long-wavelength loading. The derivation often employs auxiliary parameters to simplify computations, such as \tilde{a} = \langle h \rangle / \langle 1/a \rangle, where a = C_{44}/(C_{33} - C_{44}) represents a normalized shear compliance ratio and h is the layer's fractional thickness; similar forms are used for other parameters to relate the effective moduli. This method has been foundational for seismic interpretation, enabling the upscaling of well-log data to reservoir-scale anisotropy parameters.

Thomsen Parameters and Wave Velocities

In the context of weakly anisotropic transversely isotropic media with a vertical axis of symmetry (VTI), the Thomsen parameters offer a practical set of five dimensionless quantities to describe elastic properties relevant to seismic wave propagation. These parameters are V_P, the vertical P-wave velocity; V_S, the vertical S-wave velocity; \epsilon, the P-wave anisotropy parameter; \delta, the parameter governing near-vertical P- and SV-wave anisotropy; and \gamma, the SH-wave anisotropy parameter. They are defined in terms of the density-normalized stiffness constants C_{ij} as follows: V_P = \sqrt{C_{33}/\rho}, V_S = \sqrt{C_{44}/\rho}, \epsilon = \frac{C_{11} - C_{33}}{2 C_{33}}, \gamma = \frac{C_{66} - C_{44}}{2 C_{44}}, and \delta = \frac{(C_{13} + C_{44})^2 - (C_{33} - C_{44})^2}{2 C_{33} (C_{33} - C_{44})}. These parameters facilitate simplified expressions for phase velocities under the weak anisotropy approximation, which assumes small deviations from . The quasi-P wave velocity is approximated as V_{QP}(\theta) \approx V_P \left(1 + \delta \sin^2\theta \cos^2\theta + \epsilon \sin^4\theta \right), where \theta is the angle from the symmetry axis. The quasi-SV wave velocity follows V_{QSV}(\theta) \approx V_S \left[1 + \frac{1}{2} (\epsilon - \delta) \left( \frac{V_P}{V_S} \right)^2 \sin^2\theta \cos^2\theta \right], while the SH wave velocity is exactly V_{SH}(\theta) = V_S \sqrt{1 + 2\gamma \sin^2\theta}. The weak anisotropy approximation is valid when \epsilon, \delta, and \gamma are small, typically less than 0.1, a regime commonly observed in sedimentary basins due to fine layering or aligned minerals. This parameterization was introduced by Leon Thomsen in 1986 to simplify the analysis of anisotropic effects in seismic data. In geophysical applications, the Thomsen parameters enable seismic inversion workflows to estimate subsurface rock properties, such as fracture orientation and fluid content, by fitting observed travel times and amplitudes to these velocity models.

References

  1. [1]
  2. [2]
    General solutions for elasticity of transversely isotropic materials ...
    This article presents a short review of the harmonic general solutions for uncoupled elasticity of transversely isotropic materials with thermal and other ...
  3. [3]
    Hooke's Law for Transversely Isotropic Materials - eFunda
    Such materials are called transverse isotropic, and they are described by 5 independent elastic constants, instead of 9 for fully orthotropic. Examples of ...Missing: science | Show results with:science
  4. [4]
    Transverse Isotropy - an overview | ScienceDirect Topics
    In the context of bone, it is characterized by differences in stiffness and strength based on the direction of the applied load, with most bone tissues ...
  5. [5]
    75-plus years of anisotropy in exploration and reservoir seismics: A ...
    Jul 13, 2017 · Kelvin was also the first to formulate the elastic-wave equation for anisotropic media. (He solved it for a simple case.) Since this ...
  6. [6]
    [PDF] Materials and Elasticity Lecture M17: Engineering Elastic Constants
    Transversely Isotropic. 5 Independent elastic constants (2 E's, 2 n's, 1 G). Cubic. 3 Independent elastic constants. Isotropic. 2 Independent elastic constants ...
  7. [7]
    [PDF] Some aspects of seismic anisotropy
    Hexagonal symmetry, or, equivalently, cylindrical symmetry, is of particular interest to seismology because it is relatively ... (transverse isotropy) the ...
  8. [8]
    [PDF] Invariant-based approach to symmetry class detection - HAL
    Nov 3, 2011 · transverse isotropy ([O(2)]), tetragonal ([D4]), trigonal ([D3]) ... states that the symmetry group of an elasticity tensor is the intersection of ...
  9. [9]
    Purdue University
    Ex) Hexagonal symmetry has 5 independent elastic constants. For example, “transverse isotropy is common in seismology and has symmetry in a plane ...
  10. [10]
    Estimation of transversely isotropic formation parameters using ...
    Jul 12, 2021 · Many finely layered sedimentary formations can be described as a transversely isotropic medium with a vertical axis of symmetry (VTI).
  11. [11]
    Inherent transversely isotropic elastic parameters of over ...
    The shale samples used in this study display transverse isotropy in deformation in static and ultrasonic tests (table 1). In static tests, transverse isotropy ...3 Experimental Program · 4 Results · 4.1 Ultrasonic (dynamic)...
  12. [12]
    Effective Elastic Properties of Rocks With Transversely Isotropic ...
    Jan 5, 2019 · A theoretical model is developed to study effects of background anisotropy on elastic properties of fractured rocks · Background anisotropy has ...
  13. [13]
    [PDF] 1 CHAPTER 13 WOOD 13.1 Directionality (isotropy) 13.2 ...
    Because wood is essentially transversely isotropic, equations governing stress and strain are different from isotropic materials. The most accurate way to ...
  14. [14]
    [PDF] Temperature-dependent elasticity of single crystalline graphite - HAL
    Mar 3, 2023 · Graphite single crystal elasticity is known to possess a transverse isotropy symmetry,. 141 which is an elastic symmetry derived from the ...
  15. [15]
    Seismic properties and anisotropy of the continental crust ...
    Mar 22, 2017 · Transverse isotropy is readily found in middle and lower crustal ... A vertically transverse isotropic medium (VTI) usually shows positive radial ...
  16. [16]
    [PDF] Unified Micromechanics of Damping for Unidirectional Fiber ...
    Because of the transverse isotropy of the fibers, and consequently of the composite material: the transverse damping capacit.y of the ply along the material ...
  17. [17]
    Electrical response during indentation of piezoelectric materials
    Jan 1, 1999 · The polarization axis is the z axis, which is also the axis of transverse isotropy and the loading axis. The stress equilibrium equations ...
  18. [18]
    [PDF] Modelling non-symmetric collagen fibre dispersion in arterial walls
    Apr 15, 2015 · in-plane dispersion, transverse isotropy and isotropy. We have ... 2008 Collagen in arterial walls: biomechanical aspects. In Collagen ...
  19. [19]
  20. [20]
    Modeling Transversely Isotropic, Viscoelastic, Incompressible ... - NIH
    Such a model is useful when applying ultrasound based shear wave elastography in fibrous soft tissues such as skeletal muscles.
  21. [21]
    Constitutive laws - 3.2 Linear Elasticity - Applied Mechanics of Solids
    For an isotropic material, the elastic stress-strain relations, the elasticity matrices and thermal expansion coefficient are unaffected by basis changes. 3.2.
  22. [22]
    Matrix of Material Properties of Linear Elastic Materials
    . Such orthogonal matrix Q is termed a material symmetry. For example, if a material response is symmetric around a plane, then a reflection matrix Q_1\in ...
  23. [23]
    [PDF] On the elastic moduli and compliances of transversely isotropic and ...
    The results are applied to calculate the different sets of elastic constants for a transversely isotropic monocrystalline zinc and an orthotropic human femur. 2 ...
  24. [24]
    Modelling transversely isotropic fiber-reinforced composites with ...
    May 16, 2019 · This paper develops a micropolar constitutive model for a transversely isotropic composite material comprised of a polymer matrix and ...
  25. [25]
    Physically motivated invariant formulation for transversely isotropic ...
    This article discusses an invariant formulation for transversely isotropic hyperelasticity. The work is motivated by the interest of modeling materials such ...
  26. [26]
    (PDF) Inherent transversely isotropic elastic parameters of over ...
    Aug 6, 2025 · Experimental setup for the ultrasonic wave measurements (jacketed samples are subject to isotropic confining stress). …
  27. [27]
    [PDF] An In-Depth Tutorial on Constitutive Equations for Elastic Anisotropic ...
    Dec 1, 2011 · ... Anisotropic Elasticity Tensor. The Quarterly. Journal of Mechanics and Applied Mathematics, Vol. 42, Part 2, May 1989, pp. 249-266. Donaldson ...
  28. [28]
  29. [29]
    A robust experimental determination of Thomsen's δ parameter
    A novel inversion method for the laboratory determination of Thomsen's δ δ anisotropy parameter on cylindrical rock specimens from ultrasonic data