Resonant ultrasound spectroscopy

Last updated

Resonant ultrasound spectroscopy (RUS) is a laboratory technique used in geology and material science to measure fundamental material properties involving elasticity. This technique relies on the fact that solid objects have natural frequencies at which they vibrate when mechanically excited. The natural frequency depends on the elasticity, size, and shape of the object—RUS exploits this property of solids to determine the elastic tensor of the material. The great advantage of this technique is that the entire elastic tensor is obtained from a single crystal sample in a single rapid measurement. [1] At lower or more general frequencies, this method is known as acoustic resonance spectroscopy.

Contents

History

Interest in elastic properties made its entrance with 17th century philosophers, but the real theory of elasticity, indicating that the elastic constants of a material could be obtained by measuring sound velocities in that material, was summarized only two hundred of years later. In 1964, D. B. Frasier and R. C. LeCraw used the solution calculated in 1880 by G. Lamè and H. Lamb to solve the forward problem, and then inverted it graphically, in what may be the first RUS measurement. Nevertheless, we had to wait for the participation of geophysics community, interested in determining the Earth's interior structure, to solve the inverse problem: in 1970 three geophysicists improved the previous method and introduced the term resonant sphere technique (RST). Excited by the encouraging results achieved with lunar samples, one of them gave one of his students the task of extending the method for use with cube shaped samples. This method, now known as the rectangular parallelepiped resonance (RPR) method, was further extended by I. Ohno in 1976. Finally, at the end of the 1980s, A. Migliori and J. Maynard expanded the limits of the technique in terms of loading and low-level electronic measurements, and with W. Visscher brought the computer algorithms to their current state, introducing the final term resonant ultrasound spectroscopy (RUS). [2]

Theory

Firstly, one must solve the problem of calculating the natural frequencies in terms of sample dimensions, mass, and a set of hypothetical elastic constants (the forward problem). Then one must apply a nonlinear inversion algorithm to find the elastic constants from the measured natural frequencies (the inverse problem).

Lagrangian minimization

All RUS measurements are performed on samples that are free vibrators. Because a complete analytical solution for the free vibrations of solids does not exist, one must rely upon approximations. Finite element methods are based on balancing the forces applied to a differential volume element, then calculating its response. Energy minimization methods on the other hand determine the minimum energy, and thus the equilibrium configuration for the object. Among the energy minimization techniques, the Lagrangian minimization is the most used in the RUS analyses because of its advantage in speed (an order of magnitude smaller than the finite-element methods).

The procedure begins with an object of volume V, bounded by its free surface S. The Lagrangian is given by

where KE is the kinetic energy density

and PE is the potential energy density

Fig. 1: Computer-generated illustrations of some normal modes of vibrations for a rectangular parallelepiped sample. RUS 1.jpg
Fig. 1: Computer-generated illustrations of some normal modes of vibrations for a rectangular parallelepiped sample.

Here, is the ith component of the displacement vector, ω is the angular frequency from harmonic time dependence, is a component of the elastic stiffness tensor, and ρ is the density. Subscripts i, j, etc., refer to Cartesian coordinate directions.

To find the minimum of the Lagrangian, calculate the differential of L as a function of u, the arbitrary variation of u in V and on S. This gives:

Because is arbitrary in V and on S, both terms in square brackets must be zero. [3] Setting the first term equal to zero yields the elastic wave equation. The second square bracketed term is an expression of free surface boundary conditions; is the unit vector normal to S. For a free body (as we assume it), the latter term sums to zero and can be ignored.

Thus the set of that satisfies the previously mentioned conditions are those displacements that correspond to ω being a normal mode frequency of the system. This suggests that the normal vibrations of an object (Fig. 1) may be calculated by applying a variational method (in our case the Rayleigh-Ritz variational method, explained in the next paragraph) to determine both the normal mode frequencies and the description of the physical oscillations. [4] To quote Visscher, getting both equations from the basic Lagrangian is "a mathematical fortuity that may have occurred during a lapse in Murphy's vigilance". [5]

Rayleigh-Ritz variational method

The actuation of this approach requires the expansion of the in a set of basis functions appropriate to the geometry of the body, substituting that expression into Eq. (1) and reducing the problem to that of diagonalizing a N×N matrix (eigenvalue problem). The stationary points of the Lagrangian are found by solving the eigenvalue problem resulting from Eq. (4), that is,

where an are the approximations to the motion expanded in a complete basis set, E comes from the kinetic energy term, and Γ comes from the elastic energy term. The order of the matrices is ~10^3 for good approximations.

Equation (5) determines the resonance frequencies from the elastic moduli. [3]

The inverse problem

The inverse problem of deducing the elastic constants from a measured spectrum of mechanical resonances has no analytical solution, so it needs to be solved by computational methods. For the indirect method, a starting resonant frequency spectrum, (n=1,2,...) is calculated using estimated values for the elastic constants and the known sample dimensions and density. The difference between the calculated and measured resonance frequency spectrum, (n=1,2,...) is quantified by a figure of merit function,

where (n=1,2,...) are weight coefficients reflecting the confidence on individual resonance measurements. Then, a minimization of the function F is sought by regressing the values of all the elastic constants using computer software developed for this process. [6]

Measurements

Fig. 2: Schematic of the two transducer resonant ultrasound spectroscopy set up. RUS 4.png
Fig. 2: Schematic of the two transducer resonant ultrasound spectroscopy set up.

The most common method for detecting the mechanical resonant spectrum is illustrated in Fig. 2, where a small parallelepiped-shaped sample is lightly held between two piezoelectric transducers. One transducer is used to generate an elastic wave of constant amplitude and varying frequency, whereas the other is used to detect the sample's resonance. As a frequency range is swept, a sequence of resonance peaks is recorded. The position of these peaks occurs at the natural frequencies (from which the elastic constants are determined) and the quality factor Q (a measure of how narrow the resonance is) provides information about the dissipation of elastic energy.

Unlike in a conventional ultrasonic measure, in a method that resonates the sample a strong coupling between the transducer and the sample is not required, because the sample behaves as a natural amplifier. [2] Rather, keeping at minimum the couple between them, you get a good approximation to free surface boundary conditions and tend to preserve the Q, too. For variable-temperature measurements the sample is held between the ends of two buffer rods that link the sample to the transducers (Fig. 3) because the transducers must be kept at room temperature. In terms of pressure, on the contrary, there is a limit of only a few bars, because the application of higher pressures leads to dampening of the resonances of the sample. [1]

Samples

RUS can be applied to a great range of samples sizes, with a minimum in the order or a few hundred micrometers, but for the measurement of mineral elasticity it is used on samples typically between 1 mm and 1 cm in size.

The sample, either a fully dense polycrystalline aggregate or a single crystal, is machined in to a regular shape. [1] Theoretically any sample shape can be used, but you obtain a substantial saving in computational time using rectangular parallelepiped resonators (RPR), spherical or cylindrical ones (less time savings with cylinders).

Fig. 3: The sample assembly for a RUS variable-temperature measurement. RUS 5.jpg
Fig. 3: The sample assembly for a RUS variable-temperature measurement.

Since the accuracy of the measure depends strictly on the accuracy in the sample preparation, several precautions are taken: RPRs are prepared with the edges parallel to crystallographic directions; for cylinders only the axis can be matched to sample symmetry. RUS is rarely used for samples of lower symmetry, and for isotropic samples, alignment is irrelevant. For the higher symmetries, it is convenient to have different lengths edges to prevent a redundant resonance.

Measurements on single crystals require orientation of the sample crystallographic axes with the edges of the RPR, to neglect the orientation computation and deal only with elastic moduli. [4]

Polycrystalline samples should ideally be fully dense, free of cracks and without preferential orientation of the grains. Single crystal samples must be free of internal defects such as twin walls. The surfaces of all samples must be polished flat and opposite faces should be parallel. Once prepared, the density must be measured accurately as it scales the entire set of elastic moduli. [1]

Transducers

Unlike all other ultrasonic techniques, RUS ultrasonic transducers are designed to make dry point contact with the sample. This is due to the requirement for free surface boundary conditions for the computation of elastic moduli from frequencies. For RPRs this requires a very light touch between the sample's corners and the transducers. Corners are used because they provide elastically weak coupling, reducing loading, and because they are never vibrational node points. Sufficiently weak contact ensures no transduced correction is required. [4]

Applications

As a versatile tool for characterizing elastic properties of solid materials, RUS has found applications in a variety of fields. In geosciences one of the most important applications is related to the determination of seismic velocities in the Earth's interior. In a recent work, [7] for example, the elastic constants of hydrous forsterite were measured up to 14.1 GPa at ambient temperature. This study showed that aggregate bulk and shear moduli of hydrous forsterite increase with pressure at a greater rate than those of the corresponding anhydrous phase. This implies that at ambient conditions VP and VS of hydrous forsterite are slower than those of anhydrous one; conversely, with increasing pressure, and consequently depth, VP and VS of hydrous forsterite exceed those of anhydrous one. In addition hydration decreases the VP/VS ratio of forsterite, the maximum compressional wave azimuthal anisotropy and the maximum shear wave splitting. These data help us to constrain Earth's mantle composition and distinguish regions of hydrogen enrichment from regions of high temperature or partial melt.

Related Research Articles

<span class="mw-page-title-main">Resonance</span> Tendency to oscillate at certain frequencies

Resonance is a phenomenon that occurs when an object or system is subjected to an external force or vibration that matches its natural frequency. When this happens, the object or system absorbs energy from the external force and starts vibrating with a larger amplitude. Resonance can occur in various systems, such as mechanical, electrical, or acoustic systems, and it is often desirable in certain applications, such as musical instruments or radio receivers. However, resonance can also be detrimental, leading to excessive vibrations or even structural failure in some cases.

<i>Q</i> factor Parameter describing the longevity of energy in a resonator relative to its resonant frequency

In physics and engineering, the quality factor or Q factor is a dimensionless parameter that describes how underdamped an oscillator or resonator is. It is defined as the ratio of the initial energy stored in the resonator to the energy lost in one radian of the cycle of oscillation. Q factor is alternatively defined as the ratio of a resonator's centre frequency to its bandwidth when subject to an oscillating driving force. These two definitions give numerically similar, but not identical, results. Higher Q indicates a lower rate of energy loss and the oscillations die out more slowly. A pendulum suspended from a high-quality bearing, oscillating in air, has a high Q, while a pendulum immersed in oil has a low one. Resonators with high quality factors have low damping, so that they ring or vibrate longer.

Dynamic mechanical analysis is a technique used to study and characterize materials. It is most useful for studying the viscoelastic behavior of polymers. A sinusoidal stress is applied and the strain in the material is measured, allowing one to determine the complex modulus. The temperature of the sample or the frequency of the stress are often varied, leading to variations in the complex modulus; this approach can be used to locate the glass transition temperature of the material, as well as to identify transitions corresponding to other molecular motions.

A quartz crystal microbalance (QCM) measures a mass variation per unit area by measuring the change in frequency of a quartz crystal resonator. The resonance is disturbed by the addition or removal of a small mass due to oxide growth/decay or film deposition at the surface of the acoustic resonator. The QCM can be used under vacuum, in gas phase and more recently in liquid environments. It is useful for monitoring the rate of deposition in thin-film deposition systems under vacuum. In liquid, it is highly effective at determining the affinity of molecules to surfaces functionalized with recognition sites. Larger entities such as viruses or polymers are investigated as well. QCM has also been used to investigate interactions between biomolecules. Frequency measurements are easily made to high precision ; hence, it is easy to measure mass densities down to a level of below 1 μg/cm2. In addition to measuring the frequency, the dissipation factor is often measured to help analysis. The dissipation factor is the inverse quality factor of the resonance, Q−1 = w/fr ; it quantifies the damping in the system and is related to the sample's viscoelastic properties.

<span class="mw-page-title-main">Electron paramagnetic resonance</span> Technique to study materials that have unpaired electrons

Electron paramagnetic resonance (EPR) or electron spin resonance (ESR) spectroscopy is a method for studying materials that have unpaired electrons. The basic concepts of EPR are analogous to those of nuclear magnetic resonance (NMR), but the spins excited are those of the electrons instead of the atomic nuclei. EPR spectroscopy is particularly useful for studying metal complexes and organic radicals. EPR was first observed in Kazan State University by Soviet physicist Yevgeny Zavoisky in 1944, and was developed independently at the same time by Brebis Bleaney at the University of Oxford.

<span class="mw-page-title-main">Magnetic force microscope</span>

Magnetic force microscopy (MFM) is a variety of atomic force microscopy, in which a sharp magnetized tip scans a magnetic sample; the tip-sample magnetic interactions are detected and used to reconstruct the magnetic structure of the sample surface. Many kinds of magnetic interactions are measured by MFM, including magnetic dipole–dipole interaction. MFM scanning often uses non-contact AFM (NC-AFM) mode.

<span class="mw-page-title-main">Mechanical resonance</span> Tendency of a mechanical system

Mechanical resonance is the tendency of a mechanical system to respond at greater amplitude when the frequency of its oscillations matches the system's natural frequency of vibration closer than it does other frequencies. It may cause violent swaying motions and potentially catastrophic failure in improperly constructed structures including bridges, buildings and airplanes. This is a phenomenon known as resonance disaster.

Elastic energy is the mechanical potential energy stored in the configuration of a material or physical system as it is subjected to elastic deformation by work performed upon it. Elastic energy occurs when objects are impermanently compressed, stretched or generally deformed in any manner. Elasticity theory primarily develops formalisms for the mechanics of solid bodies and materials. The elastic potential energy equation is used in calculations of positions of mechanical equilibrium. The energy is potential as it will be converted into other forms of energy, such as kinetic energy and sound energy, when the object is allowed to return to its original shape (reformation) by its elasticity.

Nanoindentation, also called instrumented indentation testing, is a variety of indentation hardness tests applied to small volumes. Indentation is perhaps the most commonly applied means of testing the mechanical properties of materials. The nanoindentation technique was developed in the mid-1970s to measure the hardness of small volumes of material.

A mechanical amplifier, or a mechanical amplifying element, is a linkage mechanism that amplifies the magnitude of mechanical quantities such as force, displacement, velocity, acceleration and torque in linear and rotational systems. In some applications, mechanical amplification induced by nature or unintentional oversights in man-made designs can be disastrous, causing situations such as the 1940 Tacoma Narrows Bridge collapse. When employed appropriately, it can help to magnify small mechanical signals for practical applications.

Compressed sensing is a signal processing technique for efficiently acquiring and reconstructing a signal, by finding solutions to underdetermined linear systems. This is based on the principle that, through optimization, the sparsity of a signal can be exploited to recover it from far fewer samples than required by the Nyquist–Shannon sampling theorem. There are two conditions under which recovery is possible. The first one is sparsity, which requires the signal to be sparse in some domain. The second one is incoherence, which is applied through the isometric property, which is sufficient for sparse signals.

The impulse excitation technique (IET) is a non-destructive material characterization technique to determine the elastic properties and internal friction of a material of interest. It measures the resonant frequencies in order to calculate the Young's modulus, shear modulus, Poisson's ratio and internal friction of predefined shapes like rectangular bars, cylindrical rods and disc shaped samples. The measurements can be performed at room temperature or at elevated temperatures under different atmospheres.

Microrheology is a technique used to measure the rheological properties of a medium, such as microviscosity, via the measurement of the trajectory of a flow tracer. It is a new way of doing rheology, traditionally done using a rheometer. There are two types of microrheology: passive microrheology and active microrheology. Passive microrheology uses inherent thermal energy to move the tracers, whereas active microrheology uses externally applied forces, such as from a magnetic field or an optical tweezer, to do so. Microrheology can be further differentiated into 1- and 2-particle methods.

Saturated absorption spectroscopy measures the transition frequency of an atom or molecule between its ground state and an excited state. In saturated absorption spectroscopy, two counter-propagating, overlapped laser beams are sent through a sample of atomic gas. One of the beams stimulates photon emission in excited atoms or molecules when the laser's frequency matches the transition frequency. By changing the laser frequency until these extra photons appear, one can find the exact transition frequency. This method enables precise measurements at room temperature because it is insensitive to doppler broadening. Absorption spectroscopy measures the doppler-broadened transition, so the atoms must be cooled to millikelvin temperatures to achieve the same sensitivity as saturated absorption spectroscopy.

In mathematics and electronics, Cavity perturbation theory describes methods for derivation of perturbation formulae for performance changes of a cavity resonator.

The acoustoelastic effect is how the sound velocities of an elastic material change if subjected to an initial static stress field. This is a non-linear effect of the constitutive relation between mechanical stress and finite strain in a material of continuous mass. In classical linear elasticity theory small deformations of most elastic materials can be described by a linear relation between the applied stress and the resulting strain. This relationship is commonly known as the generalised Hooke's law. The linear elastic theory involves second order elastic constants and yields constant longitudinal and shear sound velocities in an elastic material, not affected by an applied stress. The acoustoelastic effect on the other hand include higher order expansion of the constitutive relation between the applied stress and resulting strain, which yields longitudinal and shear sound velocities dependent of the stress state of the material. In the limit of an unstressed material the sound velocities of the linear elastic theory are reproduced.

<span class="mw-page-title-main">Objective stress rate</span>

In continuum mechanics, objective stress rates are time derivatives of stress that do not depend on the frame of reference. Many constitutive equations are designed in the form of a relation between a stress-rate and a strain-rate. The mechanical response of a material should not depend on the frame of reference. In other words, material constitutive equations should be frame-indifferent (objective). If the stress and strain measures are material quantities then objectivity is automatically satisfied. However, if the quantities are spatial, then the objectivity of the stress-rate is not guaranteed even if the strain-rate is objective.

<span class="mw-page-title-main">Non-contact atomic force microscopy</span>

Non-contact atomic force microscopy (nc-AFM), also known as dynamic force microscopy (DFM), is a mode of atomic force microscopy, which itself is a type of scanning probe microscopy. In nc-AFM a sharp probe is moved close to the surface under study, the probe is then raster scanned across the surface, the image is then constructed from the force interactions during the scan. The probe is connected to a resonator, usually a silicon cantilever or a quartz crystal resonator. During measurements the sensor is driven so that it oscillates. The force interactions are measured either by measuring the change in amplitude of the oscillation at a constant frequency just off resonance or by measuring the change in resonant frequency directly using a feedback circuit to always drive the sensor on resonance.

<span class="mw-page-title-main">Loop-gap resonator</span>

A loop-gap resonator (LGR) is an electromagnetic resonator that operates in the radio and microwave frequency ranges. The simplest LGRs are made from a conducting tube with a narrow slit cut along its length. The LGR dimensions are typically much smaller than the free-space wavelength of the electromagnetic fields at the resonant frequency. Therefore, relatively compact LGRs can be designed to operate at frequencies that are too low to be accessed using, for example, cavity resonators. These structures can have very sharp resonances making them useful for electron spin resonance (ESR) experiments, and precision measurements of electromagnetic material properties.

Anelasticity is a property of materials that describes their behaviour when undergoing deformation. Its formal definition does not include the physical or atomistic mechanisms but still interprets the anelastic behaviour as a manifestation of internal relaxation processes. It is a special case of elastic behaviour.

References

  1. 1 2 3 4 Angel, R. J.; Jackson, J. M.; Reichmann, H. J.; Speziale, S. (2009). "Elasticity measurements on minerals: A review". European Journal of Mineralogy. 21 (3): 525. Bibcode:2009EJMin..21..525A. CiteSeerX   10.1.1.500.3003 . doi:10.1127/0935-1221/2009/0021-1925.
  2. 1 2 Maynard, J. (1996). "Resonant Ultrasound Spectroscopy". Physics Today. 49 (1): 26–31. Bibcode:1996PhT....49a..26M. doi:10.1063/1.881483.
  3. 1 2 Migliori, A.; Maynard, J. D. (2005). "Implementation of a modern resonant ultrasound spectroscopy system for the measurement of the elastic moduli of small solid specimens". Review of Scientific Instruments. 76 (12): 121301–121301–7. Bibcode:2005RScI...76l1301M. doi: 10.1063/1.2140494 .
  4. 1 2 3 Levy, Moistes; Bass, Henry E.; Stern, Richard. Celotta, Robert; Lucatorto, Thomas (eds.). Modern acoustical techniques for the measurement of mechanical properties. Experimental Methods in the Physical Sciences. San Diego: Academic Press. ISBN   978-0-12-475986-2.
  5. Visscher, W. M.; Migliori, A.; Bell, T. M.; Reinert, R. A. (1991). "On the normal modes of free vibration of inhomogeneous and anisotropic elastic objects". The Journal of the Acoustical Society of America. 90 (4): 2154. Bibcode:1991ASAJ...90.2154V. doi: 10.1121/1.401643 .
  6. Schwarz, R. B.; Vuorinen, J. F. (2000). "Resonant ultrasound spectroscopy: Applications, current status and limitations". Journal of Alloys and Compounds. 310 (1–2): 243–250. doi:10.1016/S0925-8388(00)00925-7.
  7. Mao, Z.; Jacobsen, S. D.; Jiang, F.; Smyth, J. R.; Holl, C. M.; Frost, D. J.; Duffy, T. S. (2010). "Velocity crossover between hydrous and anhydrous forsterite at high pressures". Earth and Planetary Science Letters. 293 (3–4): 250. Bibcode:2010E&PSL.293..250M. doi:10.1016/j.epsl.2010.02.025.