Skip to main content
The Journal of the Acoustical Society of America logoLink to The Journal of the Acoustical Society of America
. 2023 Dec 1;154(6):3580–3594. doi: 10.1121/10.0022581

Quantifying uniaxial prestress and waveguide effects on dynamic elastography estimates for a cylindrical rod

Melika Salehabadi 1, Lara Nammari 1, Aime Luna 1, Joseph Crutison 1, Dieter Klatt 1, Thomas J Royston 1,a),
PMCID: PMC10693442  PMID: 38038614

Abstract

Dynamic elastography attempts to reconstruct quantitative maps of the viscoelastic properties of materials by noninvasively measuring mechanical wave motion in them. The target motion is typically transversely-polarized relative to the wave propagation direction, such as bulk shear wave motion. In addition to neglecting waveguide effects caused by small lengths in one dimension or more, many reconstruction strategies also ignore nonzero, non-isotropic static preloads. Significant anisotropic prestress is inherent to the functional role of some biological materials of interest, which also are small in size relative to shear wavelengths in one or more dimensions. A cylindrically shaped polymer structure with isotropic material properties is statically elongated along its axis while its response to circumferentially-, axially-, and radially-polarized vibratory excitation is measured using optical or magnetic resonance elastography. Computational finite element simulations augment and aid in the interpretation of experimental measurements. We examine the interplay between uniaxial prestress and waveguide effects. A coordinate transformation approach previously used to simplify the reconstruction of un-prestressed transversely isotropic material properties based on elastography measurements is adapted with partial success to estimate material viscoelastic properties and prestress conditions without requiring advanced knowledge of either.

I. INTRODUCTION

Dynamic elastography methods—using noninvasive optical, ultrasonic, or magnetic resonance imaging modalities—aim to quantitatively map the shear viscoelastic properties of materials. These properties in biological tissues are often altered by disease and injury. When considering larger (relative to wavelength) regions of interest, like the liver or brain, boundary effects often can be ignored. However, as elastography expands to other anatomical regions where dimensions in at least one direction are smaller or of comparable length to bulk shear wavelengths—such as in slender skeletal muscles, blood vessels, the heart wall, and the cornea—boundary effects become non-negligible and must be considered. Researchers using optical elastography to assess the viscoelastic properties of the cornea have long recognized this, adapting models to include waveguides by treating the cornea as a plate-like structure. Here, transverse wave motion on the cornea is modeled as Rayleigh-Lamb waves.1 Blood vessels, as well, have been modeled using cylindrical shell equations considering fluid-structure interaction.2–4 Limited studies on cardiac elastography have also acknowledged the frequency-dependent (i.e., wavelength-dependent) waveguide behavior of the heart wall.5

Often, when elastography studies are done under varying nonzero quasi-static prestress conditions, observed changes in mechanical wave behavior are attributed solely to the nonlinear property of the tissue: it is observed that its shear and viscous constants are highly dependent on the static load and associated deformation. A recent article provides a summary of the literature relevant to this issue, in particular for uniaxially prestressed cylindrically-shaped structures, as well as biaxially prestressed plate-like structures.6 In another recent article, the impact of neglecting the prestress effect in cornea elastography is quantified.7

In the present study, we focus on a uniaxially prestressed cylindrical structure as an idealized geometry for skeletal muscles found throughout the body that have been studied using elastography including, muscles of the lower leg,8–12 upper leg,13–16 forearm,17 and upper arm.18–20 We investigate the confounding effects of finite dimensions and prestress on elastography measurements. We articulate and evaluate a strategy for decoupling prestress and waveguide effects from estimates of material shear viscoelastic properties based on circumferentially-, axially-, and transversely-polarized wave motion in the cylinder.

II. THEORY

A. Mechanical wave motion in a uniaxially prestressed linear viscoelastic material

Most dynamic elastography methods assume that the measured transverse wave speed or wavelength for small amplitude (linear theory assumption) motion is directly related to the material's elastic or viscoelastic properties.21 Assuming isotropy, homogeneity, and neglecting boundary effects or variation in density ρ in a nearly incompressible viscoelastic material, the frequency-dependent shear wave phase speed cω for harmonic excitation at circular frequency, ω, is

cω=ω/Realkshω=1/ρReal1/μω. (1)

Here, kshω is the complex-valued, frequency-dependent shear wave number, and μω is the complex-valued, frequency-dependent shear modulus, comprised of the shear storage modulus, μRω, and the shear loss modulus, μIω, such that μω=μRω+jμIω where j=1. The attenuation rate of the wave as it propagates is governed by the imaginary part of the wavenumber: Imagksh. In a viscoelastic material, both the shear storage and loss moduli affect both the phase speed and attenuation rate. In a purely elastic material, μI=0, there is no attenuation and the phase speed is independent of frequency (nondispersive) and reduces to c=μ/ρ. While some linear studies have assumed pure elasticity (no viscosity), often their analyses are generalizable to the linear viscoelastic problem for harmonic motion by simply adding the imaginary shear loss modulus to form the complex shear modulus. This approach is used in the present analysis.

Consider the introduction of a uniaxial static prestress σ parallel to the z-axis (Fig. 1). If the static deformation due to the prestress is assumed sufficiently small such that higher order nonlinear terms can be neglected, the governing equations expressed in polar coordinates r,φ,z are as follows: where u,v,w denote incremental displacements in the r,φ,z directions, Srr,SφφandSzz are normal stresses in the respective directions, and Srφ,Sφz,Szr,Srz,SzφandSφr are shear stresses with the first subscripted dimension referring to the shear surface and the second subscripted dimension denoting the shear direction. Finally, subscripted r,φ,z,t after a comma refer to partial derivatives with respect to that spatial or time dimension,22

Srr,r+1rSrrSφφ+Sφr,φ+Szr,z=ρu,tt, (2)
Srφ,r+1rSφr+Srφ+Sφφ,φ+Szφ,z=ρv,tt, (3)
Srz,r+1rSrz+Sφz,φ+Szz,z=ρw,tt. (4)

The normal and shear stress terms are as follows:

Srr=λ+2μu,r+λrv,φ+u+λ+λσσw,z, (5)
Sφφ=λu,r+λ+2μ1rv,φ+u+λ+λσσw,z, (6)
Szz=λ+λσσu,r+λ+λσσrv,φ+u+λ+2μ+σ+4μσσ+2λσσw,z, (7)
Srφ=μru,φv+μv,r, (8)
Sφz=μ+μσσv,z+μ+μσσ1rw,φ, (9)
Szr=μ+μσσw,r+μ+σ+μσσu,z, (10)
Srz=μ+μσσu,z+μ+μσσw,r, (11)
Szφ=μ+μσσ1rw,φ+μ+σ+μσσv,z, (12)
Sφr=μv,r+μru,φv. (13)

Here, λ is the volume elasticity of the material, a Lame constant. It is closely related to the material's bulk modulus κ=λ+(2/3)μ. In biological soft tissue or other “nearly incompressible” materials, λ and κ are multiple orders of magnitude greater than μ such that the Poisson's ratio for the material approaches, but does not equal 0.5. The terms μσ and λσ are coefficients that regulate material behavior in the presence of an initial stress, which in this case is uniaxial prestress σ aligned with the z (cylinder) axis, as shown in Fig. 1. Specifically, they account for a linear dependence of μ and λ on σ. In the Refs. 22 and 23, these terms are β1 and β2, and β1 is related to A used in other references,24–26 where its value in soft tissue-like nearly incompressible materials can range between negative and positive values. It has been hypothesized that its value depends on microstructure and thus can reveal material changes not captured by μ.24 As noted, the previous formulation neglects higher (3rd, 4th, etc.) order nonlinear terms in the strain energy function and the limitations of this simplification will be evident in the numerical and experimental studies of Secs. III and IV. Incorporation of the higher order terms, which have been detailed in previous studies investigating bulk wave motion without waveguide effects,23 is left for future study. Such an incorporation has been successfully detailed in other elastography-focused studies, in particular, to understand the impact of compressive uniaxial loading.24–26

FIG. 1.

FIG. 1.

(Color online) Uniaxially prestressed cylinder with different harmonic excitation configurations. Shaded field of view regions FOVφ,FOVz,andFOVx show the location of images presented in the following figures for torsionally-, axially-, and transversely-polarized wave motion. These are sagittal slices in the x–z or y–z plane. Deeper shaded blocks with vertical thick blue two-way arrows denote cross section of band surrounding the cylinder that delivers (z) axially-polarized harmonic actuation (b). Torsionally-polarized (a) and transversely-polarized (c) motion is input by appropriate motion at the base of the phantom, indicated by arrows. Axially-polarized (d) and transversely-polarized (e) line segment sources are also indicated at the geometric center of the phantom.

In the present study, we will assume harmonic steady state motion. By using the separation of variables one form of the solution for u,v,andw, the symmetric form, is

ur,φ,z,t=Urcosnφejωtξz, (14)
vr,φ,z,t=Vrsinnφejωtξz, (15)
wr,φ,z,t=Wrcosnφejωtξz. (16)

The anti-symmetric form is

ur,φ,z,t=Ursinnφejωtξz, (17)
vr,φ,z,t=Vrcosnφejωtξz, (18)
wr,φ,z,t=Wrsinnφejωtξz. (19)

In the following subsections, we consider specific cases of torsionally-, axially-, and transversely-polarized harmonic wave excitation that leads to simplifications of the previous expressions.

B. Torsionally-polarized wave front

Following Section 8.2.2 of Graff,27 consider waves that are torsionally-polarized and axisymmetric, meaning no variation with φ. The anti-symmetric Eqs. (17)–(19) reduce to the case of n=0, and thus ur,φ,z,t=wr,φ,z,t=0. Furthermore, we can write

vr,z,t=Vrejωtξz=Arejωtξz,AJ1βrejωtξz,β=0,β>0, (20)

where J1 denotes a Bessel function of the first kind of order 1. The radius of the cylinder will determine allowable values of β>0, for which there will be an infinite number such that, in general,

vr,z,t=A0rejωtξz+n=1AnJ1βnrejωtξnz. (21)

At =R, the cylinder boundary, Srθ=(μ/r)u,θv+μv,r=0 which means that v,r(1/r)v=0. This is satisfied for the case that β=0 and also imposes that βRJ0βR=J1βR, which leads to values β1R=5.136,β2R=8.417,β3R=11.62, While there are an infinite number of solution sets, given attenuation due to viscosity we expect the lowest wavenumber values, associated with the longest wavelengths, to dominate as we move further from the source of excitation. Additionally, the geometric configuration of the source of the excitation may preferentially drive certain wavenumber solutions.

With this form for vr,z,t, Eq. (3) reduces to the following:

(22).

μV,rr1rV,r+2μrV,r1rVξ2μ+σ+μσσV=ω2ρV, (22a)
β2+ξ21+σμ1+μσ=ksh2. (22b)

Here, ξ and β denote complex wavenumbers, with ξ in the z-direction, and β in the r-direction. Equation (22) shows how they are related to the shear wave number ksh. For the case that β=0 and σ=0 we have ξ0=ksh. In other words, the torsional waves propagate along the cylinder axis exactly as bulk shear waves, with wavelength λsh equal to 2π/Realksh. If there is an axial prestress σ then the wavelength will equal

λξ0=2πRealξ0=2πRealkshReal1+σμ1+μσ. (23)

Assuming the material to be nearly incompressible, axial prestress σ and axial prestrain ϵ are related by σ=3μ0ϵ, where μ0 is the static (real number) value of μω at ω=0. If one is able to acquire estimates of ξ0 based on fitting ejξ0z to the measured wave profile along the cylinder axis z at a given randφ at both unstressed and prestressed conditions at multiple frequencies, then μωandσ can be independently determined, even without being able to directly measure σ under static load conditions. (A direct measurement of σ may not be possible in vivo or in any situation where the cylindrical structure is connected at its ends to another structure and it is impossible to measure the stress condition between them.) Specifically, by selectively driving (or measuring via appropriate filtering) only the β=0 mode, measurements at multiple frequencies without prestress enable a determination of μω, including (possibly by extrapolation with an assumed rheological model) μ0=μω=0. Then, we can use the fact that σ=3μ0ϵ and Eq. (23) to independently determine values for σandμσ, assuming we can estimate ϵ from imaging. Numerical and experimental studies in the Secs. III and IV, respectively, illustrate and evaluate this process.

C. Axially-polarized axisymmetric wave front

We next consider the case of an axisymmetric axially-polarized wave front, which theoretically could be driven by axial oscillation of the rigid cuff encircling the cylinder over a finite axial length. This again results in n = 0, but now with the symmetric form, Eqs. (14)–(16), which reduce to

ur,z,t=Urejωtξz, (24)
wr,z,t=Wrejωtξz. (25)

Following Graff,27 Section 8.2.2, the boundary condition at the outer radius R means that Srz=Srz=0, which is satisfied by the following forms for Ur and Wr:

Ur=jξJ1βr+AαJ1αr, (26)
Wr=βJ0βr+AjξJ0αr. (27)

Here again, there will be certain allowable values of α1,α2,,α, β1,β2,,β, and ξ1,ξ2,,ξ. Inserting these forms for uandw into Eq. (4) results in the following, where Bessel functions with α and β are separated into two separate equations, Eqs. (28) and (29), that both must be satisfied,

1+μσσμβ2+1+σμ1+3μσ+λσξ2=ω2ρμ=ksh2, (28)
1+σλ+2μλσ+2μσα2+1+σλ+2μ1+4μσ+2λσξ2=ω2ρλ+2μ=kp2. (29)

Equation (28) shows the coupled relationship between axial wavenumber ξ, radial wavenumber β, and shear wavenumber ksh. Equation (29) relates ξ, another axial wavenumber α, and compression (longitudinal) wavenumber kp. If prestress σ=0, the previous equations reduce to expressions found in Graff.27

Unlike in the torsional case, even with selective excitation of, say, the lowest values for α1, β1, and ξ1, there is not a case with α or β=0. Nonetheless, Eqs. (28) and (29) do show the relationship between shear, compression, axial, and radial wavenumbers, and how this is effected by uniaxial prestress. Focusing on Eq. (28), if values for β1 and ξ1 can be determined at multiple frequencies with no prestress present, they can be used to then estimate μω. Then, prestress is added to determine σ=3μ0ϵ based on measuring ϵ. Following that, Eq. (28) can be used at different prestress levels to estimate μσ and λσ. Numerical and experimental studies in the following Secs. III and IV, respectively, illustrate and evaluate this process.

D. Transversely-polarized axisymmetric wave front

For wave symmetric to the x-axis in Fig. 1, we restrict ourselves to the case of n = 1, and Eqs. (14)–(16) reduce to the following:

ur,φ,z,t=Urcosφejωtξz, (30)
vr,φ,z,t=Vrsinφejωtξz, (31)
wr,φ,z,t=Wrcosφejωtξz. (32)

Unfortunately, this case does not simplify to the same degree as did the torsionally- and axially-polarized cases. All three displacements u, v, and w, are present, and the resulting forms for Ur, Vr, and Wr remain more complex.27 Further analytical development of these solution forms in the presence of uniaxial prestress is left for future study. In the present article, we instead consider simplifying theories that may yield insight as the cylinder's radius becomes small as compared to the wavelengths being considered. We specifically review “thin” and “thick” beam theories under uniaxial prestress conditions.

1. Prestressed one-dimensional “thin” waveguide under transverse excitation

By assuming that the radius R of the cylinder is small enough such that βR1, then so-called one-dimensional beam theory approximations can be applied. The simplest of these is the Euler–Bernoulli thin beam theory, which allows for the incorporation of prestress σ. Referring again to Fig. 1, the pre-tensioned Euler-Bernoulli thin beam theory described in Section 3.3.4 of Graff27 for x-polarized transverse wave propagation of the beam along its z-axis is the following:

EIu,zzzzσAu,zz+Aρu,tt=0. (33)

Here, I is the area moment of inertia about the y-axis [I=(π/4)R4 for a circular cross section of radius R], and A is the cross-sectional area in the x–y plane. The general solution form is: v=Vej(ωtξz) where ξ has four possible solutions,

ξ=+α,+jβ,α=γ+γ2+ω2a21/21/2,β=γ+γ2+ω2a21/21/2,γ=323R2σμ,a=R83μρ. (34)

We have two propagating waves (α) in the + or –z-direction, and two non-propagating (near field or evanescent) waves (β) in the + or –z direction. For the propagating waves, the phase speed will be: cph=(ω/Realα). Taking the limit that σ/μ 3R2/32 we see that α=ω/a1/2, and thus for the elastic case, cph=(ω1/2/2)(3μR2/4ρ)1/4, which is the classic thin (Euler–Bernoulli) beam transverse vibration solution. On the other hand, taking the limit of tension σ/μ 3R2/32 we then drop EIu,zzzz from Eq. (33) and reformulate the solution to find that there are two propagating solutions with cph=σ/ρ1/2. This is the classic transverse thin string vibration solution. A case between the extremes of either neglecting σA or EI still does not match the phase speed of bulk shear waves, given by Eq. (1).

2. Prestressed one-dimensional “thick” waveguide under transverse excitation

The thin beam formulation presented in Sec. II D 1 is only a reasonable approximation when the wavelength of the propagating transverse waves is at least an order of magnitude greater than the beam's cross-sectional radius R, or equivalent radius for a noncircular cross section. A formulation yielding a valid approximation for shorter transverse waves is based on the Timoshenko beam theory, which allows for shear deformation and accounts for rotational inertia. Incorporating prestress into the formulation for Timoshenko beam theory given in Section 3.4 of Graff,27 we have the following:

EIu,zzzzσAu,zz+Aρu,ttIρ1+Eχμu,zztt+ρ2Iχμu,tttt=0. (35)

Here, χ is the Timoshenko shear coefficient, equal to 10/9 for a circular cross section. Consider harmonic motion at frequency ω and the general solution form is u=Uej(ωtξx), and ξ has four possible solutions,

ξ4EI+ξ2σAω2Iρ1+Eκμω2ρAω2ρ2Iκμ=0,ξ=+γ+γ2+ω2a21/21/2,γ=8μ0D2μfϵLω237ρ60μ,a=3μ/ρ16D2ω29ρ10μ. (36)

In this case, we cannot separate the wavenumber solutions into propagating and non-propagating pairs. Rather, there will be certain prestress-dependent frequency ranges where we have only one propagating pair and others where we have two propagating pairs.

E. Accounting for prestress using “transformation acousto-elastography” (TAE)

Equations (23), (28), and (29) “hint” at an alternative strategy to solve the inverse problem of independently identifying inherent material shear viscoelastic properties and prestress based on measurements of wave motion. In previous studies on transversely isotropic materials not under prestress,28–32 the last author has shown that, by distorting the geometry based on direction and polarization-dependent planar phase speeds one can then solve an equivalent isotropic problem. This approach to the anisotropic problem is “Transformation Elastography.” Uniaxial prestress causes a similar, though not identical, direction and polarization dependence of the planar shear wave phase speed. The same approach is adapted to the acoustoelastic problem here, which also has the added complexity of waveguide effects, and is evaluated in numerical and experimental studies noted in the following.

The analysis in the previous sections shows that the case of torsionally-polarized waves may be the simplest and is a good place to start. Considering that the material is nearly incompressible, the prestress results in a static strain of the cylindrical phantom of unstressed length L and radius R, changing its axial length to L1+ϵ=L(1+(σ/3μ0)) and its radius to R/1+ϵ=R/1+(σ/3μ0). By driving or measuring only β=0 torsional wave motion, if one then distorts the axial length by dividing it by Real1+(σ/μ)1+μσ=Real1+(3μ0ϵ/μ[ω])1+μσ, then fits a solution to ejξ0*z in order to identify ξ0*, where an asterisk * is added to denote this is based on a distorted geometry, then we see that ξ0*=ksh, where ksh was obtained under the unstressed condition. From the degree of distortion needed to achieve this equality, we directly determine μσ.

Note, μω ( μ0,μα,α) needs to be determined for the no prestrain ( ϵ=0) case for multiple frequencies. So, instead of adjusting the geometry distortion by varying μσ as we propose here to match with the unstressed case, another approach is to adjust μσ to match the measured ξ0 in the undistorted geometry using Eq. (23). This will lead to the same result as arrived at using the geometry distortion (TAE approach). However, we propose that the TAE concept, applicable to more than the torsional polarization case, may help conceptually as an analysis tool.

Next, consider the axisymmetric, axially-polarized wavefront. In this case, the geometry is distorted by dividing the axial length by Real[1+(3μ0ε/μ[ω])(1+3μσ+λσ)] and dividing the diameter or radius by Real[1+μσ(3μ0ε/μ[ω])], such that Eq. (28) simplifies to

β*2+ξ*2=ksh2. (37)

Assume that, within the region bounded by the rigid cuff, we have purely radially converging axially-polarized waves that can be fit to J0β*r in order to solve for β*. This follows from Eq. (27) since it is assumed that ξ*=0. Additionally, with this assumption, we see from Eq. (37) that β*=ksh. Thus, like in the torsionally-polarized case, we can match β* to the value of ksh that was obtained with no prestress by adjusting μσ. This becomes another means of identifying μσ.

Then, by measurement of the wave field generated in both the radial and axial directions with and without axial prestress present, we can use Eqs. (26)–(29) to estimate λσ. However, there are an infinite number of solutions to this transcendental set of equations. Numerical and experimental studies in Sections III and IV evaluate whether this approach can be used to determine λσ. Furthermore, while a similar analysis has not been completed for the case of transversely-polarized (flexural) waves, numerical studies detailed in the following provide insight into the relationship between prestress and μσ and λσ in this case.

III. NUMERICAL CASE STUDY

A. Methods

An analytical and numerical case study was conducted to understand the interactions between uniaxial prestress and waveguide behavior, as well as to evaluate the TAE approach introduced in Sec. II E. Wave motion in the cylindrical phantom shown in Fig. 1, with parameter values provided in Table I, was simulated with a numerical finite element (FE) approach using ANSYS Mechanical APDL Version 2022 R1 (Ansys, Canonsburg, PA). These parameter values were chosen to be similar (though not exactly matching) to those in experimental studies described in Sec. IV. For simulation of torsionally- and axially-polarized harmonic excitation, an axisymmetric mixed u-P formulation was used with Plane183 8-node elements with individual element side dimensions of 0.5 mm. (Both displacements and hydrostatic pressure are taken as primary unknowns in the mixed u-P formulation, which is recommended for nearly incompressible materials.) For transverse x-polarized excitation [referring to Fig. 1(c)] a mixed u-P formulation was used with Solid273 8-node by six circumferential plane generalized axisymmetric elements with individual element side dimensions in the axial and radial direction of 0.5 mm.

TABLE I.

Parameter values for case studies.

Parameter Symbol Value(s)
Bulk Modulus κ 2  GPa
Static shear storage modulus μ0 27kPa
Spring pot parameter values μα 1.5kPa·sα
α 0.3
Undeformed phantom length L 100 mm
Undeformed phantom radius R 10, 17.5 mm
Uniaxial tensile strain ϵL 0, 0.025, 0.05, 0.1, 0.2
Density ρ 1070  kg/m3
Gent model limiting parameter Jm 50
Best fit prestress coefficient (FEM) μσ 0.75
Best fit prestress coefficient (FEM) λσ 2.00

In all cases, first, the response to the static preload is solved by accounting for geometric nonlinearity (nlgeom,on). One end of the cylinder was fixed in the axial (z)-direction and the other end was incrementally displaced in the z-direction, solving the problem in steps until the desired end displacement was reached that resulted in a uniform uniaxial strain throughout the model. This was done using the Gent model33 defined by, κ,μ0,andJm, or using linear elastic properties, E,ν,andμ0, where Young's modulus E and Poisson's ratio ν were chosen to be consistent with the value of κ used for the Gent model. For axial strains up to 20%, differences between the Gent and linear model, as quantified by different predicted axial stress values, were about half the percentage of the strain. Specifically, the axial stress predicted by the Gent model exceeded the stress predicted by the linear model by 1.25% for 2.5% axial prestrain, 2.54% for 5% prestrain, and 11.2% for 20% prestrain. While the Gent model is more accurate, it was necessary to use the linear model in Ansys for the type of linear harmonic perturbation analysis conducted here, which required that the shear modulus value be changed between the static and harmonic analysis phases. This was needed due to the frequency-dependence of the shear modulus.

Specifically, once the static analysis was done, the solution routine was exited and then re-entered, recovering the prestrained geometry and a “total” stiffness matrix that is the sum of the initial stiffness matrix due to material properties plus terms added due to the prestress condition.34 Then, the shear and Young's modulus values are updated, which in turn updates the “total” stiffness matrix. This is done before turning geometric nonlinearity off (nlgeom,off) and entering the harmonic solution routine using this modified total stiffness matrix acquired at the end of the static solution routine. Now, harmonic φ,z,orx-direction displacements were applied. For torsional and transverse (flexural) excitation, one end of the phantom was specified to have φ-polarized or x-polarixed motion, respectively, in order to preferentially drive the β=0 torsional mode or n=1 flexural mode, respectively. For axially polarized excitation, nodes on the outer surface of the phantom representing a rigid axially-oscillating cuff in contact with the phantom from 35 to 50 mm axially were given harmonic z-polarized displacement inputs. In addition to these simulations, we also computed the response to an axially- and to a transversely-polarized 10 mm long line segment displacement located at the geometric center of the cylinder, as an approximation of dynamic elastography using modulated radiation force of ultrasound to create a force vector at the focal region of the ultrasound probe.

For the harmonic analysis, a “Fractional Voigt” rheological model is assumed35 and the shear modulus is modified from μ0 to μω=μ0+μαωαcosπα/2 and viscous (beta) damping is specified by taking the ratio of the imaginary to the real part of the shear modulus and dividing by ω, the harmonic circular driving frequency. Specifically, the beta damping value is

beta=μαωαsinπα/2ωμ0+μαωαcosπα/2. (38)

B. Results

The FEA-calculated in-phase steady-state torsionally ( φ)-, axially (z)-, and transversely (x)-polarized response over a sagittal slice (FOVs in Fig. 1) is shown in Figs. 2, 3, and 4, respectively, at 0%, 10%, and 20% axial strain levels at 600 Hz for the 35 mm diameter phantom.

FIG. 2.

FIG. 2.

(Color online) Sagittal view of torsionally ( φ)-polarized in-phase wave motion from FE simulation. (a)–(c) 0%, 10%, and 20% axial prestrain, respectively.

FIG. 3.

FIG. 3.

(Color online) Sagittal view of axially (z)-polarized in-phase wave motion from FE simulation. (a)–(c) 0%, 10%, and 20% axial prestrain, respectively.

FIG. 4.

FIG. 4.

(Color online) Sagittal view of transversely (x)-polarized in-phase wave motion from FE simulation. (a)–(c) 0%, 10%, and 20% axial prestrain, respectively.

The acoustoelastography challenge is to estimate the complex frequency-dependent shear modulus and the prestress and prestress-dependent coefficients, μσandλσ, based only on assuming a known density ρ and using the complex wave images at multiple frequencies and prestress levels, from which we can also determine axial prestrain ϵ. The procedures described in Sec. II E are used to estimate the rheological properties first at multiple frequencies under zero prestrain in order to obtain estimates of μ0,μα,andα, and then with prestrain, in order to estimate prestress σ, and coefficients μσandλσ. We start with a curve fit estimate of the wave profile along the z-direction during torsionally-polarized harmonic excitation, followed by fits of the wave profile along r- and z-directions during axially-polarized harmonic excitation. Transversely-polarized (x-direction) and line segment force studies following this are used to identify the effect of prestress σ on flexural wave motion, and to assess the combined prestress and waveguide effect on shear wave motion generated from a focused source, as an approximation of a modulated radiation force of ultrasound source. Results at 600 Hz are summarized in Tables II and III for the 35 and 20 mm diameter phantoms, respectively.

TABLE II.

Wavenumber estimates for 35 mm phantom at 600 Hz at indicated prestrain levels.

Prestrain ϵL% 0 2.5 5 10 20
True Shear modulus, wavenumber μω=μR+jμIkPa=42.8+j8.05,kshm1=588j54.9
ξ0(m−1) – torsional (FEM) 588-j54.9 584-j54.6 582-j54.3 577-j53.8 571-j53.3
ksh=ξ0* (m−1) – torsional (FEM) 588-j54.9 588-j55.0 588-j54.9 590–j55.0 596-j55.6
β (m−1) – axial (FEM) 584-j60.5 596-j62.0 607-j63.5 632-j66.8 686-j74.3
ksh=β*(m−1) – axial (FEM) 584-j60.5 586-j60.9 586-j61.3 588-j62.1 587-j63.6
ξ2+β2 (m−1) – axial (FEM) 614-j56.4 615-j60.7 618-j64.6 613-j70.4 630-j74.2
ξ*2+β*2(m−1) – axial (FEM) 614-j56.4 613-j60.5 614-j64.2 599-j68.8 586-j69.0
ξ1(m−1) – flexural (FEM) 603-j68.6 594–j66.9 585-j65.2 568-j62.1 539-j56.8
ξ1* (m−1) – flexural (FEM) 603-j68.6 604-j68.0 604-j67.3 605-j64.1 607-j64.0
ξ1(m−1) – flexural (“Thin” Theory) 197-j9.18 197-j9.14 197-j9.11 196-j9.04 195-j8.89
ξ1 (m−1) – flexural (“Thick” Theory) 565-j51.5 564-j51.5 564-j51.4 564-j51.4 563-j51.3
ξ(m−1) – trans line source (FEM) 575-j179 569-j177 564-j173 552-j161 526-j136
ksh=ξa*(m−1) (FEM) 575-j179 579-j180 582-j179 564-j165 548-j142
ksh=ξb*(m−1) (FEM) 575-j179 585-j182 594-j182 587-j171 590-j153
β(m−1) – axial line source (FEM) 615-j474 627-j468 641-j464 671-j458 743-j474
ksh=β*(m−1) (FEM) 615-j474 618-j461 622-j450 629-j429 648-j413
ξ0(m−1) – torsional (Experiment) 583-j54.4 569-j40.1 555-j40.2 540-j52.2 493-j54.4
ksh=ξ0* (m−1) μσ=0.5 583-j54.4 575-j40.5 568-j41.1 565-j54.6 541-j69.7
β (m−1) – axial (Experiment) 572-j84.3 NA 586-j38.7 603-j86.0 NA
ksh=β*(m−1) (Experiment) 572-j84.3 NA 573-j37.8 576-j82.1 NA

TABLE III.

Wavenumber estimates for 20 mm phantom at 600 Hz at indicated prestrain levels.

Prestrain ϵL% 0 2.5 5 10 20
Shear modulus & wavenumber μω=μR+jμIkPa=42.8+j8.05,kshm1=588j54.9
ξ0(m−1) – torsional (FEM) 588-j54.8 585-j54.5 582-j54.2 577–j53.8 571-j53.2
ksh=ξ0* (m−1) – torsional(FEM) 588-j54.8 588-j54.8 588-j54.8 590-j55.0 596-j55.5
β (m−1) – axial (FEM) 586-j60.6 598-j61.8 610-j62.9 636-j65.4 694-j73.6
ksh=β*(m−1) – axial (FEM) 586-j60.6 588-j58.7 589-j56.6 592-j51.9 593-j41.8
ξ2+β2 (m−1) – axial (FEM) 546-j37.5 542-j34.3 536-j25.9 557-j33.8 558-j15.6
ξ*2+β*2(m−1) – axial (FEM) 546-j37.5 546-j36.8 540-j32.5 567-j41.0 578-j40.7
ξ1(m−1) – flexural (FEM) 554-j90.4 543-j80.6 534-j72.1 519-j58.3 498-j33.8
ξ1* (m−1) – flexural (FEM) 554-j90.4 552-j83.7 551-j77.9 553-j58.4 562-j49.5
ξ (m−1) – flexural (“Thin” Theory) 261-j12.1 260-j12.1 260-j12.0 259-j11.8 256-j11.5
ξ (m−1) – flexural (“Thick” Theory) 577-j50.7 576-j50.6 575-j50.5 574-j50.3 571-j50.0
ξ(m−1) – trans line source (FEM) 552-j123 536-j116 520-j109 492-j98.4 448-j83.4
ksh=ξa*(m−1) (FEM) 552-j123 539-j117 526-j111 503-j103 467-j90.9
ksh=ξb*(m−1) (FEM) 522-j123 545-j119 537-j116 524-j111 504-j104
β(m−1) – axial line source (FEM) 621-j933 627-j937 632-j939 642-j938 660-j888
ksh=β*(m−1) (FEM) 621-j933 619-j919 617-j902 609-j863 590-j737
ξ0(m−1) – torsional (Experiment) 568-j110 513-j189 499-j94.5 476-j113 484-j59.6
ksh=ξ0* (m−1) μσ=0.5 568-j110 518-j194 509-j101 495-j126 525-j81.6
β (m−1) – axial (Experiment) 574-j73.5 NA 606-j80.2 611-j88.2 NA
ksh=β*(m−1) (Experiment) 574-j73.5 NA 593-j72.5 584-j72.0 NA

Curve fits were done in matlab using the “lsqcurvefit” command. Fits for the torsionally-polarized excitation were along 5 cm axial lines, averaging fits from lines spaced 0.5 mm apart from r=0.5 mm to r=R. First, with no prestrain, it was confirmed that this fitting approach conducted at multiple frequencies enabled the identification of μ0,μα,andα of the Fractional Voigt rheological model. In Fig. 5, the black dashed line shows the real (storage) part μR versus the imaginary (loss) part μI of the shear modulus as frequency is varied. The blue x's denote the estimated values of μR verus μI based on curvefitting FE simulations at 500, 600, and 700 Hz for the 35 mm diameter phantom (they lie on the black dashed line as expected). By drawing a line through these x's, the intercept with the horizontal axis is μ0. The real and imaginary values at a given frequency can then be used to determine μαandα.

FIG. 5.

FIG. 5.

(Color online) Shear loss versus shear storage modulus. The black dashed line is actual value based on Table I values for μ0,μα,andα of the Fractional Voigt rheological model. Blue ×'s are based on the torsionally-polarized curve fit to simulations at 500, 600, and 700 Hz with zero prestrain. Fitted values are also shown for prestrain of 5% without (red circle) and with (green asterisks) the TAE correction.

For the nonzero prestrain cases, the length distortion was iterated by adjusting μσ as discussed in Sec. II E until ξ0* at 2.5% prestrain matched ksh (previously obtained with 0% prestrain) minimizing the least square error. It was found that setting, μσ=0.75, gave the best fit, with a difference in predicted ksh of less than 0.1% as compared to the unstressed case, for both the 35 and 20 mm diameter phantoms, as provided in Tables II and III, respectively. This close agreement was also present for 5% prestrain, but results diverged to about 0.4% and 1.4% difference at 10% and 20% prestrain, respectively, possibly due to geometric nonlinearities that are not accounted for in the linear-based approach. Figure 5 also shows how corresponding estimates of μ change as prestrain is increased to 5% without the TAE correction (red circle) and with the TAE correction (green asterisks).

For the axially-polarized case, first a line fit was done along the radial line within the excitation zone (cuff). It was found that μσ=0.75 also gave a good fit, with a difference in predicted ksh of less than 0.4% as compared to the unstressed case, for both the 20 and 35 mm diameter phantoms tested at 2.5% and 5% prestrain. This provided confidence in the theoretical developments of Sec. II. Results diverged up to 1.2% difference at 10% and 20% prestrain.

Next, averaged axial line fits were located at 0.5 mm spacing from r=0 to r=R extending 30 to 50 mm distance from the excitation cuff (based on the unstrained geometry) in an attempt to reduce near field effects closer to the source, which may contain larger contributions from n > 1 of the α1,α2,,α, β1,β2,,β, and ξ1,ξ2,,ξ solutions. Then, we averaged fits along radial lines 0<r<R taken at axial distances from 30 to 50 mm from the source in 0.5 mm increments. With μσ already determined, the length distortion was iterated solely by adjusting λσ until at 2.5% prestrain the values of β* and ξ* satisfied Eq. (37) where the value used for ksh was based on 0% prestrain. As expected, for this case estimates of ksh under unstressed conditions σ=0 were not accurate, being 5% too high and 7% too low for the 35 and 20 mm diameter phantoms, respectively. However, by setting λσ=2.0 it was found that cases with 2.5% prestrain yielded predictions of ksh consistent with the unstressed case within 0.2%. This difference increased to about 6% at 20% prestrain.

For the transversely-polarized (flexural wave) case, fits along axial lines were located in 0.5 mm increments from r=0 to r=R extending 30 to 50 mm distance from the excitation (based on the unstrained geometry), as in the axially-polarized study, in order to estimate ξ1. The same length distortion used in the axial study, Real[1+(3μ0ε/μ[ω])(1+3μσ+λσ)], was applied here in order to estimate ξ1*. As in the axially polarized case it is not surprising that the values of ξ1 obtained under zero prestress do not align with ksh. They are influenced by waveguide effects (small diameter of the phantom relative to shear wavelength). However, as in the axially-polarized case, by setting λσ=2.0 it was found that cases with 2.5% prestrain yielded predictions of ksh consistent with the unstressed case within 0.4%. This difference increased to only about 1.5% at 20% prestrain. This suggests that the transverse waves are affected by uniaxial prestress in the same way as the axially-polarized waves, and thus may provide a more direct estimate of λσ.

Theoretical estimates of ξ using “thick” beam theory were within 10% of values found using the finite element analysis (FEA) and reduced in value with respect to increasing prestress, like in the FEA simulations, though to not the same degree observed in the simulations. As expected, thick beam theory was a better match for the 20 mm versus the 35 mm diameter phantom. Thin beam theory results were poor in both cases. This is expected since the wavelengths generated at 600 Hz were comparable to radial dimensions. They would need to be an order of magnitude greater than radial dimensions for thin beam theory to be a good approximation.

To approximate focused modulated radiation force of ultrasound excitation, harmonic 10 mm long line displacements were input at the geometric center of the cylinder in axially- and transversely-polarized directions, as noted in Fig. 1. The response along a line orthogonal to the truncated line source was then fit to determine the complex wavenumber. If we assume that the truncated line source is an infinite plane in an infinite medium (neglecting finite boundaries) and is creating a planar shear wave front the axially-polarized wave traveling in the radial direction with wavenumber β should be governed by

1+σμμσβ2=β*2=ω2ρμ=ksh2. (39)

In other words, β*=β1+(σ/μ)μσ. This is in agreement with Eq. (28) for the case that ξ=0. Values reported in Tables II and III show that, indeed, β* at 2.5% prestrain is consistent with the value of β determined at 0% prestress, but these values differ from ksh by about 5% in terms of the real part, which is inversely proportional to shear wavelength 2π/Realksh. The imaginary value for β is an order of magnitude greater than ksh. This is expected since, not only is there attenuation due to viscosity, but also due to geometric dispersion as the finite source spreads out in three dimensions, including the axial direction, as it propagates.

Likewise, if we assume that the truncated line source is an infinite plane in an infinite medium (neglecting finite boundaries) and is creating a planar shear wave front the radially-polarized wave traveling in the axial direction (direction of uniaxial prestress) with wavenumber ξ should be governed by

1+σμ1+μσξ2=ξ*2=ω2ρμ=ksh2. (40)

In other words ξa*=ξ1+(σ/μ)1+μσ. This is not in agreement with Eq. (28) for the case that β=0, where in that case ξb*=ξ1+(σ/μ)1+3μσ+λσ. Both ξa* and ξb* are provided in Tables II and III. We expect the imaginary parts of the wave numbers to be greater than that of ksh, which is 54.9 at 600 Hz, since there will be geometric spreading. However, in this case, the geometric spreading is more confined by the small diameter of the phantom relative to its axial length. Imaginary values are larger than ksh, but by not nearly as much as they were for the other line segment case. And, as expected since there is less spreading, the imaginary part of the wavenumber is less for the 20 mm versus the 35 mm diameter phantom, since there is less geometric dispersion (∼120 versus ∼180  m1). By looking at the trend of the real part of ξa* and ξb* as prestress increases, it appears that ξb* is the better estimate for ksh, suggesting that finite diameter (waveguide effects) have impacted the transverse wave motion.

IV. EXPERIMENT

A. Optical elastography methods

The experimental configuration, shown in Fig. 6, enables excitation and measurement of torsionally-polarized wave motion on the surface of the cylindrical phantom that is simultaneously prestrained along its axis. The fixture parts were designed in Solidworks (Solidworks 2021) and three-dimensionally (3D) printed using a fused filament extrusion printer (Prusa Mk3, PRUSA REF, Prague, Czech Republic) on polyethylene terephthalate glycol (PETG). The level of prestrain, introduced by hanging weight (container of water) from the lower end, was measured using a caliper.

FIG. 6.

FIG. 6.

(Color online) Experimental setup for optical elastography using SLDV. Diagram with a bottom view and photo from the bottom and side-angle view showing cylindrical polymer phantom and tensioner. Color-coded view from bottom: (a) Bore constraint cap; (b) piezo stack; (c) rotor hub; (d) image region of interest (ROI); (e) prestressed phantom; (f) tensioner inside phantom; (g) adjustable weight; (h) Polytec SLDV scan head.

A scanning laser Doppler vibrometer (SLDV; PSV-400, Polytec, Waldbronn, Germany), is used to measure motion on the phantom that is in the direction of the laser beam, as described in previous studies.36,37 The phantoms used in this study were made of Ecoflex™ 00–30 (Smooth-On, Inc., Macungie, PA) platinum-catalyzed silicone. Ecoflex silicone bases were combined in a 1 A:1B ratio and cured at room temperature in 3D-printed molds in a vacuum chamber (5305-1212, Thermo Scientific-Nalgene, Rochester, NY) to remove any bubbles and improve phantom homogeneity.

Our group has characterized the dynamic shear viscoelastic properties of Ecoflex materials under infinitesimal (0% pre-strain) conditions over a wide frequency range.35,38 Previous publications focused on Ecoflex-10TM, a similar polymer, with greater viscosity, as compared to Ecoflex-30. For harmonic excitation of both materials using small perturbations about the unstressed state, we've found that a “fractional Voigt” rheological model best describes the frequency-dependent shear storage and loss moduli of the material. The 3-parameter fractional Voigt model is comprised of a purely elastic element of strength μ0 in parallel with a fractional order springpot that is defined by α and μα such that, at frequency, ω=2πfrad/s, the shear storage μR and loss μI moduli are given by the following equations:

μR=μ0+μαωαcosπα/2, (41)
μI=μαωαsinπα/2. (42)

From Ref. 35, it was found that for Ecoflex-10, the following values describe shear viscoelastic properties over the range from 200 Hz to 7.75 kHz: μ0=13.3kPa, μα=2kPa·sα, and α=1/3. For an Ecoflex-30 sample, measurements conducted in the same way as in the Ref. 24 over the range from 200 Hz to 1 kHz yielded: μ0=27kPa, μα=1.5kPa·sα, and α=0.3. Thus, while Ecoflex-30 is “stiffer” (higher μ0) under static conditions, due to its lower viscosity the magnitude of its complex shear modulus increases at a slower rate with frequency ( α=0.3), as compared to that of Ecoflex-10 ( α=1/3). Note, while these values for Ecoflex-30 were used in the numerical finite element study of Sec. III, the torsional experimental study described in this section while confirming the appropriateness of the Fractional Voigt model, yielded different parameter values. This could be due to variations in batches of Ecoflex-30, as well as the different experimental configuration.

SLDV measurements were made over a grid in a ROI as indicated in Fig. 6. By driving the two piezoceramic stack actuators (P842.10, PI USA, Auburn, MA) in phase only torsional motion should be excited in the ideal case that the experimental setup is perfectly symmetric in all aspects, which is never the case. It is expected that some flexural (transverse) wave motion will also be excited. If measurements at equal radial distances from the central axis at the same axial (z) position, are subtracted from one another (difference) this should double the β=0 torsional mode and eliminate the n = 1 flexural mode. If the measurements are added, it should eliminate the torsional mode and double the n = 1 flexural mode. This was done before taking line profiles axially to determine ξ0(m−1) in the torsional case and ξ1(m−1) in the flexural case from the same measurement.

B. Optical elastography results

Optical elastography measurements, processed to amplify torsional motion, as described in Sec. IV A, are shown in Fig. 7 at 0%, 5%, and 10% axial prestrain for the 35 mm diameter phantom. (Image is flipped vertically relative to diagram and photo in Fig. 5.) Corresponding estimates of ξ0(m−1) in the torsional case are provided in Tables II and III for the 35 and 20 mm diameter phantoms under different prestrain conditions. Estimates of ξ1(m−1) in the flexural case were found to be unreliable indicating that the torsional motion was significantly greater in amplitude, as was intended. The case with no prestrain yielded an estimate of the shear wavenumber of ksh=ξ0*=583j54.4 (m−1) for the 35 mm phantom, and ksh=ξ0*=568j110 (m−1) for the 20 mm phantom, suggesting the 35 mm phantom material matched values used in the FEA simulation within 1%, but the 20 mm phantom was slightly stiffer by 3% and significantly more viscous than the material parameter values used in the computational (FEA) simulations, where ksh=ξ0*=588j54.9.

FIG. 7.

FIG. 7.

(Color online) Sagittal view of torsionally ( φ)-polarized in-phase wave motion from experimental study. (a)–(c) 0%, 10%, and 20% axial prestrain, respectively. Measurement differs from the FE result in Fig. 2 since the SLDV is measuring motion in the y-direction on the rounded surface of the cylindrical phantom, not at a central sagittal cross section.

A more comprehensive study of the 20 mm phantom was undertaken by conducting the analysis over the frequency range from 200 to 600 Hz under 0%, 2.5%, 5%, 10%, and 20% prestrain levels in order to fully evaluate the material parameter identification study, starting with identifying the unstressed fractional Voigt model parameters. Figure 8 shows the estimated complex shear modulus μω under 0% pre-strain (blue lines) based on the experimental estimate of ξ0ω=kshω and μω=ρω/kshω2. The black lines are the corresponding least square error curve fit assuming a fractional Voigt model for μω. The resulting values for the model are μ0=31.2kPa,μα=140kPa·sα,andα=0.613. As compared to the model used in the numerical studies, this one has similar values for the shear storage modulus, but about twice the shear loss modulus in the frequency range of interest. Here, μω/2π=600Hz=43.6+17.8kPa. Also, due to unmodeled dynamics in the experimental system, the assumptions necessary for the torsionally-polarized inverse approach start to break down at frequencies above 500 Hz, resulting in large fluctuations in the estimated complex shear modulus.

FIG. 8.

FIG. 8.

(Color online) The complex shear modulus for the 20 mm phantom based on torsional wave measurements. (a) Shear storage modulus. (b) Shear loss modulus. Key: blue line,  0% prestrain; black dashed line, least square error fit of fractional Voigt model to 0% prestrain case; red line, 10% prestrain; green line, 10% prestrain after length distortion per TAE.

Figure 8 also shows the experimental estimate of μω when a 10% prestrain is imposed. As expected, the value of μω is increased by the prestrain. Distorting the axial length as in the numerical study by dividing it by Real1+(3μ0ϵ/μω)1+μσ where μσ=0.5 results in μ*ω, denoted by the green line in Fig. 8, which more closely agrees with the value for μω obtained under the no prestrain case. Table IV summarizes experimentally fit wavenumber estimates ξ0 and TAE-adjusted estimates of ksh=ξ0*, all based on μσ=0.5 for the same prestrain levels used in the numerical study.

TABLE IV.

Wavenumber estimates for 20 mm phantom at indicated prestrain levels.

Prestrain ϵ% 0 2.5 5 10 20
ξ0(m−1) at 200 Hz 208-j25.1 199-j23.0 199-j25.5 195-j20.1 195-j24.9
ksh=ξ0* (m−1) μσ=0.5 208-j25.1 202-j23.3 205-j26.3 207-j21.2 217-j27.7
ξ0(m−1) at 300 Hz 299-j45.6 292-j45.6 291-j45.7 288-j33.8 279-j36.7
ksh=ξ0* (m−1) μσ=0.5 299-j45.6 296-j46.2 298-j46.9 303-j35.6 308-j40.5
ξ0(m−1) at 400 Hz 381-j66.6 375-j74.0 372-j61.1 368-j56.0 349-j46.1
ksh=ξ0* (m−1) μσ=0.5 381-j66.6 380-j74.9 382-j62.6 387-j58.8 383-j50.6
ξ0(m−1) at 500 Hz 484-j80.4 469-j79.9 464-j72.4 459-j82.6 419-j116
ksh=ξ0* (m−1) μσ=0.5 484-j80.4 475-j80.1 475-j74.1 480-j86.5 458-j127
ξ0(m−1) at 600 Hz 568-j110 513-j189 499-j94.5 476-j113 484-j59.6
ksh=ξ0* (m−1) μσ=0.5 568-j110 519-j191 510-j96.7 497-j118 527-j64.9

For the experimental studies, the value of μσ=0.5 provided a better fit than the value of μσ=0.75 used in the FEA studies. The value of μσ=0.5 is consistent with the initially-developed theory of acoustoelasticity articulated by Biot in his seminal text.39 More recent studies have shown that Biot's theory is a special case of a broader theory that also incorporates higher order effects.23 Experimental results in Tables II and III show that higher order terms are necessary for accuracy as prestrain increases with the estimate of ksh=ξ0* deviating by more than 6% at 20% prestrain.

C. Magnetic resonance elastography methods

Axially- and transversely-polarized wave propagation in the same uniaxially prestressed cylindrical isotropic phantoms used in the optical measurements was studied experimentally using magnetic resonance elastography. All experiments were conducted in a 30 cm horizontal bore Agilent 9.4 Tesla preclinical magnetic resonance imaging (MRI) system using a 120 mm I.D. 600 mTesla/meter maximum strength gradient set and a 72 mm I.D. linear radio frequency coil.

The experimental fixture to hold the phantom in the magnet and simultaneously apply a known static tensile preload while also delivering vibratory excitation via a rigid cuff around the phantom at its mid-length is shown in Fig. 9. The fixture parts were designed in Solidworks (Solidworks 2021) and 3D printed using a fused filament extrusion printer (Prusa Mk3, PRUSA REF) on PETG to limit interference with magnetic resonance (MR) signals and keep parts light enough to be moved efficiently by the piezo stack actuator that delivered the vibratory excitation. It is an advancement of a prior design.40 Tensile pre-loading of the cylinder is shown in green in Fig. 9. Samples are gripped by two clamp types: a fixed clamp at the proximal end, and the tensioner clamp at the distal end. Both clamp types are slotted to allow wooden skewers to penetrate the sample, providing reasonably distributed tension, rather than surface tension that would be only available with typical clamps. The distal clamp is then attached to a reinforced nylon wire that is fed out the back of the magnet bore and attached to a pulley system where adjustable weight applies appropriate tension to achieve a desired measurable strain.

FIG. 9.

FIG. 9.

(Color online) Experimental setup for magnetic resonance elastography (MRE). Diagram with top and side view, and photo from top showing cylindrical polymer phantom and claw-grip tensioner. Color-coded view from side: (a) Bore constraint; (b) piezo counter mass; (c) piezo stack; (d) fixed phantom end clamp; (e) harmonic actuator; (f) image ROI; (g) prestressed phantom; (h) claw-grip tensioner; (i) pulley (outside magnet); (j) adjustable weight.

Simultaneous harmonic actuation (orange in Fig. 9) is achieved through a piezo stack (P842.10, PI USA LP) located proximally to the entry of the bore to limit interference from wiring, and attached to a Delrin countermass to maximize transference of motion to the phantom. The phantom is placed in a cylindrical tube and the actuator is attached directly to the piezo and slotted for adjustable cuffs (yellow) that uniformly and circumferentially grip the sample, providing a fixed location of the wavesource that contacts the phantom over 15 mm in length axially. Harmonic vibration is produced along the same axis as the pre-tension. Note, however, that because of gravity and imperfections in this setup flexural (transverse) motion will also be excited.

SLIM MRE41 was used to acquire vibratory motion encoded in three orthogonal directions simultaneously. Sequence parameters were as follows: TR/TE 1600/16 ms; eight time steps were captured at even intervals over the period of the vibration frequency. A 250 mT/m motion encoding gradient (MEG), which was matched to the vibration frequency, was used with ten cycles. Measurements were repeated with inverted polarity gradients to subtract static field inhomogeneities. The data matrix size of 64 × 64 × 40 with an isotropic voxel size of 0.75 mm resulted in a field of view (FOV) of 48 mm × 48 mm × 30 mm. The piezo stack provided an input axial harmonic motion of ∼10 μm peak amplitude at 600 Hz.

D. MRE results

Sagittal views obtained using MRE of the axially-polarized vibratory motion are provided in Fig. 10 for 0%, 5%, and 10% axial prestrain for the 35 mm diameter phantom. The field of view is the same as in the numerical study of Sec. III (except shortened to 3 cm along the axis). The images were analyzed using the same TAE approach detailed in Sec. III for the numerical study, with calculation results provided in Tables II and III. Estimates of β based on the radially converging axially-polarized wavefront within the cuff yielded results similar in value to the torsional studies for the 35 mm phantom, given that ksh=β*. Like in the numerical studies, as prestrain is increased we find that the measured β value increases, as expected per equation (37). A corrected estimate of β* can be obtained by dividing the diameter by Real1+μσ(3μ0ϵ/μω). “Corrected” values β* for nonzero prestrain cases more closely match the zero prestrain case, as can be seen in the last row of Tables II and III. Corrections were made based on the experimentally-estimated rheological model found with the multi-frequency torsional study.

FIG. 10.

FIG. 10.

(Color online) Sagittal view of axially (z)-polarized in-phase wave motion from experimental study. (a)–(c) 0%, 5%, and 10% axial prestrain, respectively.

Given the limited FOV in the experimental study, it was found that estimates of ξ or ξ1 based on axially- or transversely-polarized (flexural) wave analysis were not reliable. In the finite element study, curve fits for these wavenumbers were based on wave motion 30 to 50 mm from the source, whereas in the experimental study the FOV only extended to 30 mm from the source. Signal-to-noise ratio (SNR) was too poor beyond that range.

V. DISCUSSION AND CONCLUDING REMARKS

The theoretical, numerical finite element and experimental studies of the previous sections have explored the confounding effects of finite dimensions and nonzero prestress on the elastography approach to estimating material viscoelastic properties in an isotropic cylindrical structure under uniaxial normal stress aligned with the cylinder axis. Additionally, a coordinate transformation approach—TAE—was introduced to estimate material viscoelastic properties independent of the prestress condition without requiring a priori knowledge of either the viscoelastic properties or stress conditions. Rather, only the amount of deformation, or strain, from the unstressed condition is required. Once viscoelastic properties are calculated, prestress can also be estimated.

The numerical and experimental studies show both the promise and implementation challenges of elastography and the TAE approach. In the numerical FEA study of Sec. III, the material shear storage μR and loss μI moduli, as well as the uniaxial normal stress σ could be determined with accuracy within a few percent based on torsional- or axially-polarized excitation, though accuracy generally degrades as the prestrain reaches 10% and 20%. The TAE approach formulated in Sec. II is inherently based on an assumption of linearity and as deformation increases accuracy will suffer. The approach could be improved by accounting for higher order terms in the quasi-static deformation model; this is left for future study. Related to this, in the present study it is assumed that the constitutive relation of elastic materials can be generalized to viscoelastic materials by substituting the real modulus with the complex modulus even if this linearization of the dynamic problem occurs after a finite static deformation (prestrain) has been applied. Further study is needed to evaluate this assumption.

A noted “divergence” between the FEA-based and experimental studies was the observation that the parameter μσ was best approximated as −3/4 in the FEA simulation, whereas a value of –1/2 did better in the experimental study. Recall that this term accounts for the linear dependence of μ on σ. In the Refs. 22 and 23 it is often denoted β1 and is related to A used in other references,24–26 where its value in soft tissue-like nearly incompressible materials can range between negative and positive values. It is hypothesized that its value depends on microstructure and thus can reveal material changes not captured by μ.24 Its approximate value of –1/2 in the experimental studies falls within the range found in the literature. The value of −3/4 found in the FEA studies, relative to –1/2, would lessen the effect of the prestress. This is commensurate with the fact that the linear material model used in the numerical study produces less of a change in stress levels for a given strain than does the Gent model, which has been shown to be accurate for the phantom material, Ecoflex-30.33

Whereas the unprestressed torsional FEA studies perfectly reproduced the implemented rheological model over multiple frequencies, the multi-frequency experimental study illustrated the approximate nature of the rheological model under no prestress conditions. A least square error fit over the range from 200 to 600 Hz was used to estimate μ0,μα,andα. This in turn was used in estimating the value of μσ. If the baseline (unprestressed) rheological model is a poor fit, estimates of μσ based on it would, of course, be inaccurate, as well.

Another source of error is the confounding effect of multiple wave types being present. This is particularly true for the axially-polarized and transverse-(flexural) polarized cases. Selecting a field of view away from the source and filtering can help with some of this, but not all of it. Even in the finite element studies, estimates of material shear properties based on flexural waves at any prestrain level were not accurate.

Simulations of a localized line segment source, an approximation of focused modulated radiation force of ultrasound that is commonly used in ultrasound-based elastography, highlighted how both small dimensions and prestress can alter estimates of the shear viscoelastic properties if those estimates are based on assuming bulk shear wave propagation and do not account for boundary and preloading effects.

Some “next steps” for advancing the strategy introduced here to decouple prestress and waveguide behavior from material shear stiffness estimates include consideration of more complex geometry and stress conditions, as well as anisotropic and nonuniform material properties. Finite element models based on medical images that can provide detailed geometry and localized deformation information under varying loading conditions and, if needed, measures of anisotropy and inhomogeneity, may provide a way to advance the TAE technique beyond simple geometries and assumptions of isotropy and homogeneity.42

ACKNOWLEDGMENTS

No benefits in any form have been or will be received from a commercial party related directly or indirectly to the subject of this paper. The authors acknowledge financial support from the National Science Foundation (NSF) Grant No. 1852691 and the National Institutes of Health (NIH) Grant No. AR071162.

AUTHOR DECLARATIONS

Conflict of Interest

There are no conflicts to disclose.

DATA AVAILABILITY

The raw numerical and experimental data used in this study is available from the corresponding author upon request.

References

  • 1. Larin K. V. and Sampson D. D., “ Optical coherence elastography—OCT at work in tissue biomechanics,” Biomed. Opt. Express 8(2), 1172–1202 (2017). 10.1364/BOE.8.001172 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Bernal M., Nenadic I., Urban M. W., and Greenleaf J. F., “ Material property estimation for tubes and arteries using ultrasound radiation force and analysis of propagating modes,” J. Acoust. Soc. Am. 129, 1344–1354 (2011). 10.1121/1.3533735 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Roy T. and Guddati M. N., “ Shear wave dispersion analysis of incompressible waveguides,” J. Acoust. Soc. Am. 149(2), 972–982 (2021). 10.1121/10.0003430 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Roy T., Urban M. W., Xu Y., Greenleaf J., and M. N, Guddati , “ Multimodal guided wave inversion for arterial stiffness: Ethodology and validation in phantoms,” Phys. Med. Biol. 66(11), 115020 (2021). 10.1088/1361-6560/ac01b7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Urban M. W., Qiang B., Song P., Nenadic I. Z., Chen S., and Greenleaf J. F., “ Investigation of the effects of myocardial anisotropy for shear wave elastography using impulsive force and harmonic vibration,” Phys. Med. Biol. 61(1), 365–382 (2016). 10.1088/0031-9155/61/1/365 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Crutison J., Sun M., and Royston T. J., “ The combined importance of finite dimensions, anisotropy, and pre-stress in acoustoelastography,” J. Acoust. Soc. Am. 151(4), 2403–2413 (2022). 10.1121/10.0010110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Sun M. G., Son T., Crutison J., Guaiquil V., Lin S., Nammari L., Klatt D., Yao X., Rosenblatt M. I., and Royston T. J., “ Optical coherence elastography for assessing the influence of intraocular pressure on elastic wave dispersion in the cornea,” J. Mech. Behav. Biomed. Mater. 128, 105100 (2022). 10.1016/j.jmbbm.2022.105100 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Bernabei M., Lee S. S. M., Perreault E. J., and Sandercock T. G., “ Shear wave velocity is sensitive to changes in muscle stiffness that occur independently from changes in force,” J. Appl. Physiol. 128(8), 8–16 (2020). 10.1152/japplphysiol.00112.2019 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Chernak L. A., DeWall R. J., Lee K. S., and Thelen D. G., “ Length and activation dependent variations in muscle shear wave speed,” Physiol. Meas. 34, 713–721 (2013). 10.1088/0967-3334/34/6/713 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Jenkyn T. R., Ehman R. L., and An K., “ Noninvasive muscle tension measurement using the novel technique of magnetic resonance elastography (MRE),” J. Biomech. 36, 1917–1921 (2003). 10.1016/S0021-9290(03)00005-8 [DOI] [PubMed] [Google Scholar]
  • 11. Basford J. R., Jenkyn T. R., An K. N., Ehman R. L., Heers G., and Kaufman K. R., “ Evaluation of healthy and diseased muscle with magnetic resonance elastography,” Arch. Phys. Med. Rehabil. 83, 1530–1536 (2002). 10.1053/apmr.2002.35472 [DOI] [PubMed] [Google Scholar]
  • 12. Koo T. K., Guo J.-Y., Cohen J. H., and Parker K. J., “ Quantifying the passive stretching response of human tibialis anterior muscle using shear wave elastography,” Clin. Biomech. 29, 33–39 (2014). 10.1016/j.clinbiomech.2013.11.009 [DOI] [PubMed] [Google Scholar]
  • 13. Klatt D., Papazoglou S., Braun J., and Sack I., “ Viscoelasticity-based MR elastography of skeletal muscle,” Phys. Med. Biol. 55, 6445–6459 (2010). 10.1088/0031-9155/55/21/007 [DOI] [PubMed] [Google Scholar]
  • 14. Kodesho T., Taniguchi K., Kato T., Mizoguchi S., Yamakoshi Y., Watanabe K., Fujimiya M., and Katayose M., “ Relationship between shear elastic modulus and passive force of the human rectus femoris at multiple sites: A Thiel soft-embalmed cadaver study,” J. Med. Ultrason. 48(2), 115–121 (2021). 10.1007/s10396-020-01076-w [DOI] [PubMed] [Google Scholar]
  • 15. Barnhill E., Kennedy P., Hammer S., van Beek E. J. R., Brown C., and Roberts N., “ Statistical mapping of the effect of knee extension on thigh muscle viscoelastic properties using magnetic resonance elastography,” Physiol. Meas. 34, 1675–1698 (2013). 10.1088/0967-3334/34/12/1675 [DOI] [PubMed] [Google Scholar]
  • 16. Otsuka S., Shan X., and Kawakami Y., “ Dependence of muscle and deep fascia stiffness on the contraction levels of the quadriceps: An in vivo supersonic shear-imaging study,” J. Electromyogr. Kinesiol. 45, 33–40 (2019). 10.1016/j.jelekin.2019.02.003 [DOI] [PubMed] [Google Scholar]
  • 17. Zonnino A., Smith D. R., Delgorio P. L., and Johnson C. L., “ MM-MRE: A new technique to quantify individual muscle forces during isometric tasks of the wrist using MR elastography,” in Proceedings of the 2019 IEEE 16th International Conference on Rehabilitation Robotics (ICORR), Toronto, Canada (June 24–28, 2019), pp. 270–275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. Eby S. F., Song P., Chen S., Chen Q., Greenleaf J. F., and An K. N., “ Validation of shear wave elastography in skeletal muscle,” J. Biomech. 46(14), 2381–2387 (2013). 10.1016/j.jbiomech.2013.07.033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Gennisson J. L., Cornu C., Catheline S., Fink M., and Portero P., “ Human muscle hardness assessment during incremental isometric contraction using transient elastography,” J. Biomech. 38, 1543–1550 (2005). 10.1016/j.jbiomech.2004.07.013 [DOI] [PubMed] [Google Scholar]
  • 20. Wang A. B., Perreault E. J., Royston T. J., and Lee S. S. M., “ Changes in shear wave propagation within skeletal muscle during active and passive force generation,” J. Biomech. 94, 115–122 (2019). 10.1016/j.jbiomech.2019.07.019 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Manduca A., Bayly P. J., Ehman R. L., Kolipaka A., Royston T. J., Sack I., Sinkus R., and Van Beers B. E., “ MR elastography: Principles, guidelines, and terminology,” Magn. Resonance Med. 85(5), 2377–2390 (2020). 10.1002/mrm.28627 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Kocakaplan S. and Tassoulas J. L., “ Wave propagation in initially-stressed elastic rods,” J. Sound Vib. 443, 293–309 (2019). 10.1016/j.jsv.2018.11.045 [DOI] [Google Scholar]
  • 23. Destrade M. and Ogden R. W., “ On stress-dependent elastic moduli and wave speeds,” IMA J. Appl. Math. 78(5), 965–997 (2013). 10.1093/imamat/hxs003 [DOI] [Google Scholar]
  • 24. Gennisson J. L., Renier M., Catheline S., Barriere C., Berco J., Tanter M., and Fink M., “ Acoustoelasticity in soft solids: Assessment of the nonlinear shear modulus with the acoustic radiation force,” J. Acoust. Soc. Am. 122(6), 3211–3219 (2007). 10.1121/1.2793605 [DOI] [PubMed] [Google Scholar]
  • 25. Destrade M., Gilchrist M. D., and Saccomandi G., “ Third- and fourth-order constants of incompressible soft solids and the acousto-elastic effect,” J. Acoust. Soc. Am. 127(5), 2759–2763 (2010). 10.1121/1.3372624 [DOI] [PubMed] [Google Scholar]
  • 26. Destrade M. and Ogden R. W., “ On the third- and fourth-order constants of incompressible isotropic elasticity,” J. Acoust. Soc. Am. 128(6), 3334–3343 (2010). 10.1121/1.3505102 [DOI] [PubMed] [Google Scholar]
  • 27. Graff K. F., Wave Motion in Elastic Solids ( Dover Publications Inc., New York, 1975). [Google Scholar]
  • 28. Guidetti M. and Royston T. J., “ Analytical solution for converging elliptic shear wave in a bounded transverse isotropic viscoelastic material with nonhomogeneous outer boundary,” J. Acoust. Soc. Am. 144(4), 2312–2323 (2018). 10.1121/1.5064372 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Guidetti M. and Royston T. J., “ Analytical solution for diverging elliptic shear wave in bounded and unbounded transverse isotropic viscoelastic material with nonhomogeneous inner boundary,” J. Acoust. Soc. Am. 145(1), EL59–EL65 (2019). 10.1121/1.5088028 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Guidetti M., Caratelli D., and Royston T. J., “ Converging super-elliptic torsional shear waves in a bounded transverse isotropic viscoelastic material with nonhomogeneous outer boundary,” J. Acoust. Soc. Am. 146(5), EL451–EL457 (2019). 10.1121/1.5134657 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Royston T. J., “ Analytical solution based on spatial distortion for a time-harmonic Green's function in a transverse isotropic viscoelastic solid,” J. Acoust. Soc. Am. 149(4), 2283–2291 (2021). 10.1121/10.0004133 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Guidetti M., Klatt D., and Royston T. J., “ Transformation elastography: Converting anisotropy to isotropy,” in Proceedings of the IEEE 17th International Symposium on Biomedical Imaging (ISBI), Iowa City, IA (April 3–7, 2020). [Google Scholar]
  • 33. Steck D., Qu J., Kordmahale S. B., Tscharnuter D., Muliana A., and Kameoka J., “ Mechanical responses of Ecoflex silicone rubber: Compressible and incompressible behaviors,” J. Appl. Polym. Sci. 136(5), 47025-1–47025-11 (2019). 10.1002/app.47025 [DOI] [Google Scholar]
  • 34.Ansys, Inc., “ Linear perturbation analysis,” in ANSYS Mechanical APDL Theory Reference ( Ansys, Inc., Canonsburg, PA, 2023), Chap. 15, Sec. 8. [Google Scholar]
  • 35. Yasar T. K., Royston T. J., and Magin R. L., “ Wideband MR elastography for viscoelasticity model identification,” Magn. Resonance Med. 70, 479–489 (2013). 10.1002/mrm.24495 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36. Brinker S., Kearney S. P., Royston T. J., and Klatt D, “ Simultaneous magnetic resonance and optical elastography acquisitions: Comparison of displacement images and shear modulus estimations using a single vibration source,” J. Mech. Behav. Biomed. Mater. 84, 135–144 (2018). 10.1016/j.jmbbm.2018.05.010 [DOI] [PubMed] [Google Scholar]
  • 37. Dore M., Luna A., and Royston T. J., “ Biaxial tensile prestress and waveguide effects on estimates of the complex shear modulus using optical-based dynamic elastography in plate-like soft tissue phantoms,” ASME J. Eng. Sci. Med. Diagnos. Ther. 6(1), 0110061 (2023). 10.1115/1.4056103 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Liu Y., Yasar T. K., and Royston T. J., “ Ultra wideband (500 Hz to 16 kHz) MR elastography for robust shear viscoelasticity model identification,” Phys. Med. Biol. 59, 7717–7734 (2014). 10.1088/0031-9155/59/24/7717 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Biot M. A., Mechanics of Incremental Deformation ( John Wiley & Sons, New York, 1965). [Google Scholar]
  • 40. Brinker S. and Klatt D., “ Demonstration of concurrent tensile testing and magnetic resonance elastography,” J. Mech. Behav. Biomed. Mater. 63, 232–243 (2016). 10.1016/j.jmbbm.2016.06.020 [DOI] [PubMed] [Google Scholar]
  • 41. Klatt D., Yasar T. K., Royston T. J., and Magin R. L., “ Sample interval modulation for the simultaneous acquisition of displacement vector data in magnetic resonance elastography: Theory and application,” Phys. Med. Biol. 58, 8663–8675 (2013). 10.1088/0031-9155/58/24/8663 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Crutison J. and Royston T. J., “ The design and application of a diffusion tensor informed finite element model for exploration of uniaxially prestressed muscle architecture in magnetic resonance imaging,” Eng. Comput. 38, 3893–3908 (2022). 10.1007/s00366-022-01690-x [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

The raw numerical and experimental data used in this study is available from the corresponding author upon request.


Articles from The Journal of the Acoustical Society of America are provided here courtesy of Acoustical Society of America

RESOURCES