Nonlinear derating of high-intensity focused ultrasound beams using Gaussian modal sums

Similar documents
DETERMINATION OF THE B/A OF BIOLOGICAL MEDIA BY MEASURING AND MODELING NONLINEAR DISTORTION OF PULSED ACOUSTIC WAVE IN TWO-LAYER MEDIA TAMARA KUJAWSKA

Angular Spectrum Decomposition Analysis of Second Harmonic Ultrasound Propagation and its Relation to Tissue Harmonic Imaging

Localization of Focused-Ultrasound Beams in a Tissue Phantom, Using Remote Thermocouple Arrays

Effects of Nonlinearity on the Estimation of In Situ Values of Acoustic Output Parameters

A Comparative Evaluation of Four Acoustic Hydrophones in High Intensity Focused Ultrasound (HIFU) Field Measurements

Transient, planar, nonlinear acoustical holography for reconstructing acoustic pressure and particle velocity fields a

High Intensity Focused Ultrasound Test & Measurement Techniques

CHAPTER 1 INTRODUCTION TO ACOUSTIC SATURATION

J. Tavakkoli Institute of Biomaterials and Biomedical Engineering, University of Toronto, 4 Taddle Creek Road, Toronto, Ontario M5S 3G9, Canada

APPLICATION OF THE KHOKHLOV-ZABOLOTSKAYA-KUZNETSOV EQUATION TO MODELING HIGH-INTENSITY FOCUSED ULTRASOUND BEAMS

Workshop 2: Acoustic Output Measurements

19 th INTERNATIONAL CONGRESS ON ACOUSTICS MADRID, 2-7 SEPTEMBER 2007

ABSTRACT HEAT GENERATION RATES FOR THERAPEUTIC APPLICATIONS. Esteban Echeverría Fernández, M.S., 2012

Let us consider a typical Michelson interferometer, where a broadband source is used for illumination (Fig. 1a).

Microwave-induced thermoacoustic tomography using multi-sector scanning

Ultrasonics xxx (2010) xxx xxx. Contents lists available at ScienceDirect. Ultrasonics. journal homepage:

MODEL SELECTION IN ULTRASONIC MEASUREMENTS ON TRABECULAR BONE

Simulation of nonlinear Westervelt equation for the investigation of acoustic streaming and nonlinear propagation effects

Lateral Blood Flow Velocity Estimation Based on Ultrasound Speckle Size Change With Scan Velocity

Output intensity measurement on a diagnostic ultrasound machine using a calibrated thermoacoustic sensor

Overturning of nonlinear acoustic waves in media with power-law attenuation

Nonlinear effects in ultrasound fields of diagnostic-type transducers used for kidney stone propulsion: characterization in water

Ultrasonic Measurement of Minute Displacement of Object Cyclically Actuated by Acoustic Radiation Force

Nonlinear focusing of ultrasonic waves by an axisymmetric diffraction grating embedded in water

31545 Medical Imaging systems

EXPERIMENTAL VALIDATION OF THE SPECTRAL FIT ALGORITHM USING TISSUE MIMICKING PHANTOMS

Fresnel Number Concept and Revision of some Characteristics in the Linear Theory of Focused Acoustic Beams

ULTRASONIC INSPECTION, MATERIAL NOISE AND. Mehmet Bilgen and James H. Center for NDE Iowa State University Ames, IA 50011

Physics and Knobology

PARAMETER ESTIMATION IN ULTRASONIC MEASUREMENTS ON TRABECULAR BONE

Vector diffraction theory of refraction of light by a spherical surface

The development of algebraic methods to compute

28th Seismic Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

On the beam deflection method applied to ultrasound absorption measurements

Miao Boya and An Yu Department of Physics, Tsinghua University, Beijing , People s Republic of China

Characterization Methods of High-Intensity Focused Ultrasound-Induced Thermal Field

Pixel-based Beamforming for Ultrasound Imaging

Two-Dimensional Blood Flow Velocity Estimation Using Ultrasound Speckle Pattern Dependence on Scan Direction and A-Line Acquisition Velocity

NUMERICAL METHODS FOR NONLINEAR WAVE PROPAGATION IN ULTRASOUND

10. OPTICAL COHERENCE TOMOGRAPHY

Simulation of Contrast Agent Enhanced Ultrasound Imaging based on Field II

ATTENUATION AND POROSITY ESTIMATION USING THE FREQUENCY-INDEPENDENT PARAMETER Q

Interaction of weak shocks leading to Mach stem formation in focused beams and reflections from a rigid surface: numerical modeling and experiment

Correlation analysis of three-dimensional strain imaging using ultrasound two-dimensional array transducers

EFFECTS OF ACOUSTIC SCATTERING AT ROUGH SURFACES ON THE

Simultaneous estimation of attenuation and structure parameters of aggregated red blood cells from backscatter measurements

Measurement of high intensity focused ultrasound fields by a fiber optic probe hydrophone

Metadamping: An emergent phenomenon in dissipative metamaterials

Feasibility of non-linear simulation for Field II using an angular spectrum approach

Effects of self-steepening and self-frequency shifting on short-pulse splitting in dispersive nonlinear media

Fast simulation of nonlinear radio frequency ultrasound images in inhomogeneous nonlinear media: CREANUIS

Introduction to Nonlinear Acoustics

Horns as particle velocity amplifiers

I have nothing to disclose

Plane Wave Medical Ultrasound Imaging Using Adaptive Beamforming

Estimating the spatial autocorrelation function for ultrasound scatterers in isotropic media

Acoustic field characterization of the Duolith: Measurements and modeling of a clinical shock wave therapy device

EDDY-CURRENT nondestructive testing is commonly

Technical University of Denmark

Time Domain Simulation of Harmonic Ultrasound Images and Beam Patterns in 3D Using the k-space Pseudospectral Method

Bayesian approach to image reconstruction in photoacoustic tomography

Acoustic Scattered Field Computation

Statistical Properties of Nonlinear Diffracting N-Wave behind a Random Phase Screen 1

Ultrasonic particle and cell separation and size sorting

On the Early Development of Dispersion in Flow through a Tube with Wall Reactions

Demonstration of a Novel Equation for the Near Field Distance Calculation for Annular Phased Array Transducers

PII S (98) AN ANALYSIS OF ELASTOGRAPHIC CONTRAST-TO-NOISE RATIO

A New Finite-Difference Method for General Long-Range Rotorcraft Acoustics: Initial Comparisons with Intermediate- Range Data

Lesson 05: Intensity Measurements and Bioeffects. This lesson contains 18 slides plus 6 multiple-choice questions.

A method of the forward problem for magneto-acousto-electrical tomography

Today s menu. Last lecture. Ultrasonic measurement systems. What is Ultrasound (cont d...)? What is ultrasound?

Investigation of the Energy Shielding of Kidney Stones by Cavitation Bubble Clouds during Burst Wave Lithotripsy

EL-GY 6813/BE-GY 6203 Medical Imaging, Fall 2016 Final Exam

N db compared with that in the single pulse harmonic imaging mode, whereas

INFRARED IMAGING METHOD FOR THERMOGRAPHIC CHARACTERIZING OF MEDICAL HIGH INTENSITY FOCUSED ULTRASOUND DEVICES MOHAMMADREZA JAHANGIR MOGHADAM

Research on sound absorbing mechanism and the preparation of new backing material for ultrasound transducers

ON NUMERICAL CALCULATION OF ACOUSTIC FIELD OF PLANE SOURCE

Two-photon nonlinearity in general cavity QED systems

Shear waves in solid-state materials

Numerical Simulation of the Bubble Cloud Dynamics in an Ultrasound Field

Lecture #8 Non-linear phononics

NIH Public Access Author Manuscript Ultrason Imaging. Author manuscript; available in PMC 2013 November 21.

INVESTIGATION OF ACOUSTIC STREAMING JETS IN LIQUID

Doppler echocardiography & Magnetic Resonance Imaging. Doppler echocardiography. History: - Langevin developed sonar.

The parametric propagation in underwater acoustics: experimental results

EXPERIMENTAL EVALUATION OF NONLINEAR INDICES FOR ULTRASOUND TRANSDUCER CHARACTERIZATIONS TIMOTHY ALLEN BIGELOW. B.S., Colorado State University, 1998

Exploiting pattern transformation to tune phononic band gaps in a two-dimensional granular crystal

Magnetoacoustic tomography with current injection

ECMUS The Safety Committee of EFSUMB : Tutorial

Plasma heating in stellarators at the fundamental ion cyclotron frequency

Scattering of light from quasi-homogeneous sources by quasi-homogeneous media

Full Wave Analysis of RF Signal Attenuation in a Lossy Rough Surface Cave Using a High Order Time Domain Vector Finite Element Method

Threshold of singularity formation in the semilinear wave equation

A model for the ultrasonic field radiated by an Electro-Magnetic Acoustic Transducer in a ferromagnetic solid

Physical principles of Harmonic Imaging Min Joo Choi, PhD

SLIDE NOTES. Notes on COVER PAGE

Today s menu. Last lecture. Measurement of volume flow rate. Measurement of volume flow rate (cont d...) Differential pressure flow meters

arxiv: v1 [physics.med-ph] 21 Jan 2018

Ultrasonic Testing Classroom Training Book, Second Edition

Original Contribution

Transcription:

Nonlinear derating of high-intensity focused ultrasound beams using Gaussian modal sums Seyed Ahmad Reza Dibaji and Rupak K. Banerjee Mechanical Engineering Program, School of Dynamic Systems, University of Cincinnati, 2600 Clifton Avenue, Cincinnati, Ohio 45221 Joshua E. Soneson and Matthew R. Myers a) Division of Solid and Fluid Mechanics, Center for Devices and Radiological Health, United States Food and Drug Administration, 10903 New Hampshire Avenue, Silver Spring, Maryland 20993 (Received 28 March 2013; revised 10 September 2013; accepted 23 September 2013) A method is introduced for using measurements made in water of the nonlinear acoustic pressure field produced by a high-intensity focused ultrasound transducer to compute the acoustic pressure and temperature rise in a tissue medium. The acoustic pressure harmonics generated by nonlinear propagation are represented as a sum of modes having a Gaussian functional dependence in the radial direction. While the method is derived in the context of Gaussian beams, final results are applicable to general transducer profiles. The focal acoustic pressure is obtained by solving an evolution equation in the axial variable. The nonlinear term in the evolution equation for tissue is modeled using modal amplitudes measured in water and suitably reduced using a combination of source derating (experiments in water performed at a lower source acoustic pressure than in tissue) and endpoint derating (amplitudes reduced at the target location). Numerical experiments showed that, with proper combinations of source derating and endpoint derating, direct simulations of acoustic pressure and temperature in tissue could be reproduced by derating within 5% error. Advantages of the derating approach presented include applicability over a wide range of gains, ease of computation (a single numerical quadrature is required), and readily obtained temperature estimates from the water measurements. [http://dx.doi.org/10.1121/1.4824336] PACS number(s): 43.25.Zx, 43.35.Wa, 43.35.Yb, 43.80.Ev [CCC] Pages: 3435 3445 I. INTRODUCTION In the design and testing of high-intensity focused ultrasound (HIFU) devices, as well as in treatment planning of focused ultrasound surgery, it is important to know the acoustic pressure and temperature fields in the tissue of interest. Pressure measurements are much more readily performed in water than in in vivo models, ex vivo samples, and even tissue phantoms. Techniques for converting measurements in water to estimates appropriate for tissue, i.e., derating techniques, can be valuable for characterizing HIFU systems. The simplest derating technique involves taking measured output acoustic pressure in water and reducing it by a factor based upon typical tissue attenuation, for example, 0.3 db/cm/mhz (Center for Devices and Radiological Health, 2008). This is typically applied to the center frequency of pulses used in diagnostic ultrasound applications. Derating using a single multiplicative factor can be effective for low amplitude ultrasound beams, where generation of higher harmonics due to nonlinear propagation effects is not present. During the propagation of finite-amplitude sound beams, energy is transferred between the harmonic modes due to nonlinear self-interaction; energy is also scattered and converted into heat at rates that depend on the frequency of the modes. This interplay of energy transfer and loss makes a) Author to whom correspondence should be addressed. Electronic mail: matthew.myers@fda.hhs.gov derating a highly nontrivial task. It requires knowledge of the frequency dependence of material absorption, as studied by Verma et al. (2005a, 2005b) and Rielly et al. (2000). In an attempt to account for broadband pulses or spectral broadening due to higher harmonic generation, Schafer (1990) introduced wideband derating, in which the waveform is Fourier transformed, multiplied by the appropriate frequency-dependent attenuation and corresponding phase velocity dispersion for tissue, and transformed back to the time domain. Schafer also considered other techniques, such as derating only the dominant spectral peak of the pulse, and derating the centroid of the pulse frequency spectrum. These methods were compared against data from propagation in water but not against tissue data. Szabo et al. (1999) observed that varying the power of pulses in water to match the peak negative acoustic pressure of pulses in a tissue mimicking material (TMM) resulted in close agreement of time average intensity and peak positive acoustic pressure. The results were better than the wideband derating approach of Schafer, but only useful with a priori knowledge of the TMM. Taking advantage of the strong focusing typical of therapeutic beams, Bessanova et al. (2010) scaled the transducer acoustic power, rather than the attenuation, to derate acoustic pressure data in water. This was done to match the focal acoustic pressures in water and tissue, relying on the assumption that linear acoustic pressure gain G is high enough that most of the nonlinear interaction occurs in the focal region. The derated acoustic pressure is computed J. Acoust. Soc. Am. 134 (5), November 2013 0001-4966/2013/134(5)/3435/11/$30.00 3435

using the scaled initial source amplitude. Agreement between derated water simulations and tissue simulations was reported as 35% for weakly focused diagnostic transducers (G ¼ 5 10) and 5% for strongly focused (G ¼ 40 60) therapeutic transducers. An alternative method for derating measurements involves first decomposing the measurements into Gaussian modal sums (Wu and Du, 1990; Wu and Nyborg, 1992; Soneson and Myers, 2007). Soneson and Myers (2007) showed that the acoustic pressure field in some cases can be accurately modeled using an approximate solution to the Khokhlov Zabolotskaya Kuznetsov (KZK) equation in the form of a sum of Gaussian modes. The technique is well suited to the low-gain case. Additionally, Myers and Soneson (2009) derived a method for using the derated modal amplitudes to predict the temperature rise in the tissue of interest. While the results are derived in the context of a Gaussian weighted transducer, no assumption of Gaussian shading is required in the final results. The next section demonstrates how derating is performed within the Gaussian-mode theory. Subsequently, simulated acoustic pressure measurements in water are used to obtain the modal amplitudes, which are then derated and used to construct estimates of acoustic pressure and temperature in tissue. The results are compared with direct simulation of the acoustic pressure and temperature fields in the tissue, using the full KZK equation. II. NONLINEAR DERATING MODEL We consider an acoustic pressure field periodic in time, written as pðr; z; tþ ¼ 1 2 X 1 n¼1 ½A n ðr; zþe inx 0t þ A n ðr; zþe inx 0t Š: (1) Here p is the deviation from ambient pressure, z and r are the axial and radial coordinates, x 0 is the angular frequency, and t is time measured in the reference frame propagating at the small-signal sound speed c 0. In terms of the scaled variables f ¼ z/d, q ¼ r/a,andu k ¼ A k /p 0,whered is the transducer focal length, a the effective radius, and p 0 the peak acoustic pressure deviation at the transducer face, the KZK equation (Hamilton and Morfey, 1998) governing propagation of the HIFU beam may be written as (Soneson and Myers, 2007) @u n @f þ i 4nG r2 q u n þ c n u n ¼ inn 4 X 1 m¼1 u m ðu n m þ 2u m nþ; n¼ 1; 2;::: : (2) Here G ¼ pa 2 f/(dc 0 ) is the linear acoustic pressure gain (with f ¼ x/2p), and N ¼ 2p p 0 bdf/(q 0 c 3 0 ) is the coefficient of nonlinearity, where b is the nonlinear parameter for the propagation medium and q 0 the mass density. The complex parameter c n is defined by c n ¼ daðf Þ 2id p aðf Þlog f ; g ¼ 1; (3a) f c n ¼ daðf Þ 2id p ½aðf Þ a ðf =f ÞŠ; g ¼ 2; (3b) where g is the exponent in the frequency-dependent absorption: aðf Þ¼a ðf =f Þ g : (3c) Here a is the absorption at some reference frequency f indicated by the asterisk subscript, often 1 MHz. In the computations, we will assume g¼ 1 for soft tissue and g¼2 for water. Soneson and Myers (2007) showed that it is advantageous computationally to represent the acoustic pressure field by a sum of Gaussian modes: with u n ðq; fþ ¼a n ðfþ exp ½ GnbðfÞq 2 Š; bðfþ ¼ 1 ig G ðg þ iþf : (4a) (4b) In this case the KZK equation simplifies to the following ordinary differential equation: da n df þ c G þ i n a n G ðgþiþf ¼ inn 4 X 1 m¼1 a m ða n m þ 2a m nþ: (5) Equation (5) describes the evolution of the acoustic pressure modes with axial distance. The terms in the square brackets represent the effects of absorption, phase velocity dispersion, and diffraction, while the right-hand side of the equation arises due to nonlinear propagation effects. Computing the field in this Gaussian-mode approximation is much faster than other radial discretizations. If the radial coordinate is discretized over J nodes, then solutions using the Gaussianmode method are obtained in 1/J the amount of computational time and memory required by finite difference, finite element, or spectral methods. A Green-function solution to the bioheat equation (Nyborg and Wu, 1994; Nyborg, 1988; Myers and Soneson, 2009) can be used to derive an expression for the axial temperature arising from absorption of the ultrasound field in a medium possessing the absorption characteristics given in Eqs. (3b) and (3c). The equation for the focal temperature, in a form applicable to both Gaussian and non-gaussian transducers, is (Myers and Soneson, 2009) Tð0; 1Þ ¼ a 0 4Kq 0 c 0 X 1 n¼1 n g a 2 n ð0; 1Þr2 n;eff log 1 þ 4jt! rn;eff 2 ; (6a) where K is the thermal conductivity and r n,eff is the effective radius of the nth mode. For a Gaussian transducer, r n,eff is the radius at which the beam intensity of the nth mode falls to 1/e of its on-axis value. For a non-gaussian transducer, the main lobe of the intensity mode may be fitted to a 3436 J. Acoust. Soc. Am., Vol. 134, No. 5, November 2013 Dibaji et al.: Nonlinear derating with Gaussian modes

Gaussian profile by determining the intensity at an off-axis location r, and computing r n,eff from (Myers and Soneson, 2009) r n;eff ¼ r =½lnðI n ð0; fþ=i n ðr ; fþþš 1=2 : (6b) Pressure for non-gaussian transducers may be computed from Eqs. (1) and (4), and the relationship u k ¼ A k /p 0. The result for the focal acoustic pressure is re-assembled here: p ðq ¼ 0; f ¼ 1; tþ ¼ p X 0 1 ½a n ð1þe inx0t þ a n 2 ð1þe inx 0t Š: n¼1 (7) If the acoustic pressure-mode amplitudes a n (1) and the beam widths r n,eff appropriate for tissue can be estimated from values measured in water, Eqs. (6) and (7) can be used to compute the focal acoustic pressure and temperature in a tissue medium. The derating strategy for performing the conversion from water to tissue is as follows. The right side of Eq. (5) represents the effects of nonlinear acoustic propagation. If nonlinear propagation effects in tissue can be estimated from the corresponding nonlinear effects in water, i.e., if the terms on the right side of Eq. (5) can be computed from measurements of the modal amplitudes in water, then Eq. (5) can be solved to obtain the tissue modes. To this end, we set q n ðfþ ¼ inn X 1 w m ðfþ½w n m ðfþþ2w m nðfþš; (8) 4 m¼1 where w n is an estimate for a n obtained by making measurements in water in conjunction with a derating procedure. The function q n (f) is assumed to be a known quantity. Computation of q(f) is discussed subsequently. Here we proceed with solution of the equation da n df þ h nðfþa n ¼ q n ðfþ; (9a) where G þ i h n ðfþ ¼ c n : (9b) G ðgþiþf To obtain a particular solution to (9a), we multiply both sides by expð Ð f 0 h nðuþduþ, use (9b) in the integrand, and perform the integration. A homogeneous solution to (9a) can be readily obtained after moving h n ðfþa n to the right side and dividing through by a n. The resulting total solution is a n ðfþ ¼a 1 ð0þe c G nf G ðgþiþf ð f þ e c n ðf f 0 Þ G ðgþiþf0 G ðgþiþf qðf0 Þdf 0 : (10) 0 Equation (10) provides the mechanism by which the modal amplitudes are obtained. The first term, equal to the total solution when N ¼ 0, is the solution in the limit of linear acoustics. It predicts that the focal acoustic pressure amplitude is simply the amplitude at the transducer a 1 (0), scaled by the attenuating factor exp( ad), and multiplied by the gain. In the presence of nonlinear propagation effects (N > 0), the second term in Eq. (10) contributes the modal amplitudes for the higher harmonics. The critical step in solving Eq. (10) is obtaining the approximate modal amplitudes w n contained in the source term q n ðfþ. The procedure for determining the w n values involves measurement of pressure amplitudes in water, e.g., by a hydrophone scan along the z axis. Devising the protocol for performing the water measurements, and for incorporating the water measurements in a manner that yields the best approximation to pressure measurements in tissue, is a primary objective of the nonlinear derating algorithm. One method for performing the nonlinear derating is to make water measurements using the same source pressure as in tissue, then to multiply the nth amplitude by exp( naz), where a is the tissue absorption. This is similar to the wideband derating performed by Schafer (1990). This type of derating, which we label endpoint scaling since amplitude reduction is performed after the modes have evolved from the transducer to the axial endpoint, is a logical extension of linear derating. However, since the harmonics are not attenuated during their evolution to the extent they would be in tissue, nonlinear effects can be overestimated. To address this problem, endpoint scaling is a combined with source scaling, a version of which was introduced by Bessanova et al. (2010). With source scaling, the experiments in water are performed at a lower value of the nonlinearity parameter in water than in tissue, by reducing the source acoustic pressure. That is, N w /N t < 1. When source scaling is performed, endpoint scaling by the full factor exp( naz) can result in excessive amplitude reduction. Therefore, we propose combining source scaling with endpoint scaling by the factor exp( enaz), where e 1. Numerical experiments were performed to determine the relative amounts of source scaling (represented by N w /N t ) and endpoint scaling (represented by e) that produced derated amplitudes w n giving the best agreement with numerical simulations in tissue. Agreement was measured in terms of the maximum and minimum of the acoustic pressure waveform, and the temperature rise at the focus. III. COMPUTATIONS A. Computational design Simulated acoustic pressure fields in water were obtained using the software HIFU_Simulator (Soneson, 2008), which solves the axisymmetric KZK equation for continuous wave beams using a split-step time and frequency domain algorithm on a finite difference discretization. HIFU_Simulator was also used to generate the focal acoustic pressures and temperatures in tissue against which the derating algorithm was compared. Transducer gains G of 20, 40, and 60 were considered. A representative tissue attenuation, 55 db/m at 1 MHz, was used for all simulations. Values of the coefficient of nonlinearity in tissue, N t, equal to 0.1, 0.2, and 0.3 were considered. However, results are J. Acoust. Soc. Am., Vol. 134, No. 5, November 2013 Dibaji et al.: Nonlinear derating with Gaussian modes 3437

presented only for N t ¼ 0.3, which represented the largest amount of nonlinearity and the biggest challenge to the derating algorithm. A wide range of N w /N t and e values were used in the simulations. For each value of N w /N t considered, HIFU_Simulator was executed in a water environment using a source acoustic pressure that yielded a value of N w equal to N t (0.1, 0.2, or 0.3) times the current N w /N t. After the simulation completed, the acoustic pressure amplitudes at the focus (i.e., z ¼ d or f ¼ 1) were subsequently derated by the factor exp( enad), yielding the derated water amplitudes w n. The w n values were then used in (8) and (10) to compute the estimates of the tissue amplitudes, a n. Numerical quadrature in (10) was performed using the trapezoidal rule. The modal intensities in the focal plane for water were plotted and the modal beam widths computed from (6b). A value of 0.4 mm was used for r. Once a n and r were determined, the focal acoustic pressure and temperature in tissue could be constructed from (6) and (7). The tissue acoustic pressures and temperatures determined through this derating process were compared with direct simulations in tissue performed with HIFU_Simulator. For acoustic pressure, an error metric was composed from the difference between direct simulation and derating in the peak positive acoustic pressures, added to the similar error in peak negative acoustic pressure. The error was normalized by the peak-to-peak acoustic pressure computed via direct simulation in tissue. That is, pressure error ¼ jp þ;direct p þ;derating jþjp ;direct p ;derating j : p þ;direct jp ;direct j (11a) The temperature error was the difference between the end of sonication temperature rise for direct simulation minus the end-of-sonication temperature for derating, normalized by the direct-simulation temperature rise: temperature error ¼ jdt direct DT derating j DT direct : (11b) The sonication period was that required to raise the temperature about 60 (to avoid boiling in a phantom), typically a few seconds, depending upon the value of the transducer gain. B. Derating from reduced data The function q n (f) [Eq. (8)] capturing the effects of nonlinear propagation is typically small outside of the focal region. As a result, the integrand in (10) can be negligibly small away from the focus. Hence, it is often not necessary to determine the water modal amplitudes along the entire axial distance between the source and transducer. This has important implications in practice, as the water amplitudes are determined experimentally, and a reduced axial interval of integration implies less measurement time. The axial lengths over which amplitude data was required, in order to compute acoustic pressure and temperature fields accurate to a small tolerance (less than 10% difference with direct tissue simulations), were determined computationally for different transducer gains. C. Results In Fig. 1, contours of acoustic pressure error [Eq. (11a)] are plotted as a function of e and N w /N t, for gains of 20 [Fig. 1(a)], 40 [Fig. 1(b)], and 60 [Fig. 1(c)]. The contours may be interpreted using the following example. For a gain of 40, suppose experiments are performed in water at a reduced source acoustic pressure (relative to the desired source acoustic pressure in tissue) such that with N w ¼ 0.61 N t, and the measured acoustic pressure amplitudes are reduced using the tissue absorption scaled down by a factor of 0.38. Then, the resulting derated acoustic pressure amplitudes will approximate those that would be measured directly in tissue within a 5% error. The acoustic pressure waveform in tissue, derived from the nonlinear derating process for (N w /N t, e) ¼ (0.61, 0.38), is shown in Fig. 2. The acoustic pressure waveform generated through a direct simulation in tissue is also shown. The waveform produced by the derating technique is slightly broader than that generated by direct simulation, and the peak positive and negative acoustic pressures are quite similar. As the gain increases, the bands in the contours of Fig. 1 shift toward higher values of e, implying that an increased level of endpoint reduction is required as the level of focusing increases. The error bands also become more vertical as the gain increases, meaning that a wider range of source acoustic pressures in water can accurately produce the desired acoustic pressure field in tissue, provided that the absorption value is properly selected. Contours of temperature error (Fig. 3) are very similar to those of acoustic pressure. For the example above where (N w /N t, e) ¼ (0.61, 0.38), the expected error between the measured temperature in tissue and that predicted from nonlinear derating is again less than 5%. Temperature as a function of time is plotted in Fig. 4 for this example, for both the derating approach and direct simulation of temperature rise in tissue. At the end of sonication, the difference in temperature rise predicted by the derating technique and direct tissue simulation is about 1 C. The contours in Fig. 3 are slightly less vertical than those in Fig. 1 for acoustic pressure, i.e., for a given source acoustic pressure in water (N w /N t value), a wider range of absorption values can be used in the endpoint derating to generate the correct tissue temperature to a given level of accuracy than pressure. The thickness of the contours is also less for temperature, implying that the error increases more rapidly for temperature as one move away from optimal derating conditions. The large amount of information contained in the contours of Figs. 1 and 3 can be distilled to a more practical form by considering a single value of N w /N t and plotting error as a function of the sole parameter e. We consider the case of N w /N t ¼ 0.7, i.e., derating experiments are performed in water under conditions where the coefficient of 3438 J. Acoust. Soc. Am., Vol. 134, No. 5, November 2013 Dibaji et al.: Nonlinear derating with Gaussian modes

FIG. 1. (Color online) (a) Contours of relative difference between focal acoustic pressure predicted by the derating technique and by direct simulation in tissue, as a function of the parameters N w /N t and e. G ¼ 20. Coefficient of Nonlinearity in tissue N t ¼ 0.3. (b) Contours of relative difference between focal acoustic pressure predicted by the derating technique and by direct simulation in tissue, as a function of the parameters N w /N t and e. G ¼ 40. Coefficient of Nonlinearity in tissue N t ¼ 0.3. (c) Contours of relative difference between focal acoustic pressure predicted by the derating technique and by direct simulation in tissue, as a function of the parameters N w /N t and e. G ¼ 60. Coefficient of Nonlinearity in tissue N t ¼ 0.3. nonlinearity is 70% of the value in tissue. In Fig. 5(a), the error (relative to values computed by direct simulation in tissue) for the derated acoustic pressure values is plotted as a function of e for this case. It can be seen that, for a gain of 20, the value of e that yields the minimum error for acoustic pressure is approximately 0.2. Thus, to complete the derating process (initiated by performing source scaling with N w ¼ 0.7 N t ), endpoint scaling should be performed by reducing the attenuation of each mode to 20% of its value appropriate for the tissue of interest at the given frequency. For a gain of 40, the attenuation of each mode should be reduced to approximately 42% of its non-derated value. The factor for a gain of 60 is about 61%. The minimum errors for the derated temperatures [Fig. 5(b)] occur when e ¼ 0.18 J. Acoust. Soc. Am., Vol. 134, No. 5, November 2013 Dibaji et al.: Nonlinear derating with Gaussian modes 3439

FIG. 2. (Color online) Pressure waveform computed by derating method with N w /N t ¼ 0.61, e ¼ 0.38, along with acoustic pressure waveform produced by direct simulation in tissue. G ¼ 40. Coefficient of Nonlinearity in tissue N t ¼ 0.3. (G ¼ 20), 0.42 (G ¼ 40), and 1.0 (G ¼ 60). Thus, to attain the minimum error in temperature, full attenuation should be used for each mode. For the highest gain (G ¼ 60), the optimal scaling factor e for the attenuation is considerably different for acoustic pressure and temperature. We note that it is not necessary to use the same value of e for both acoustic pressure and temperature computations. However, for simplicity, it can be useful to do so. For the gain of 60, a compromise between acoustic pressure and temperature accuracy could produce an e value around 0.7. This value yields both acoustic pressure and temperature errors less than 10%. Regarding derating using reduced data sets, computations revealed that agreement with direct simulations could be maintained with less than 10% difference when the first 80% of the data in the axial direction (0 < f < 0.8) was ignored. In Figs. 6(a) and 6(b), the focal waveform and temperature rise are shown for a case where the gain is 40 and (N w /N t, e) ¼ (0.61, 0.38). In comparison with full simulation in tissue, reduced data set derating resulted in a decrease in accuracy from 3% to 4% for acoustic pressure, and from 2% to 9% for temperature. IV. DISCUSSION The premise of the derating strategy presented in this paper is that the nonlinear coupling effects in tissue can be determined from water measurements, provided that the pressure field is reduced (more precisely, the dimensionless parameter N is reduced) at the source by the proper amount (scale factor N w /N t ) when the water experiments are conducted. This proper amount can only be determined through a large battery of numerical experiments; we performed these calculations and computed the resulting error (relative to the standard of direct simulations in tissue). If the source is not reduced, the coupling effects are overpredicted. This source reduction N w /N t alone, because it allows the complex nonlinear interactions in tissue to be computed from measurements in water, leads to reasonably accurate tissue pressures and temperatures. However, better accuracy is attained by further reducing the modal amplitudes. Since the pressures were already reduced at the source, reduction at the target location using the full attenuation is too severe. The parameter e was introduced to quantify how much of the full attenuation is required. [Each mode is attenuated a different amount, by the factor exp( eanz), as described in Sec. III A.] In terms of attaining minimum error defined by the metrics in Eqs. (11a) and (11b), the choices of N w /N t and e are not unique. That is, for any given set of experimental conditions, there are multiple choices of N w /N t and e that will yield the same errors (relative to simulations in tissue or actual measurements in tissue) in maximum and minimum pressure and maximum temperature. However, even though these errors are the same for different (N w /N t, e) combinations, other features of the pressure waveform will vary. These include RMS error over the whole waveform (which is related to the distribution of energy between harmonics), or just the peak negative pressure. So if a user happened to be particularly interested in accurately predicting peak negative pressure, say for determining the likelihood of cavitation, the user would want to select from the multiple (N w /N t, e) combinations the one that most accurately predicted peak negative pressure. Further numerical experiments would be required to identify this particular combination, as error in peak negative pressure was not a metric that was thoroughly explored in our study. In practice, the error contours in Figs. 1 and 3 would be used to transform pressure measurements in water to focal pressure and temperature in tissue using the following sequence of steps. We illustrate using the specific example of a 1.98 MHz, 5.0 cm diameter transducer with a 6.26 cm focal length. The transducer acoustic power in tissue is approximately 42 W. (1) The transducer gain G and nonlinear parameter N t are computed, using the transducer properties, the desired source acoustic pressure, and properties of the tissue medium. For the present case of interest, the gain G ¼ 40 and N t ¼ 0.3. (Acoustic pressure at transducer surface ¼ 0.34 MPa.) (2) We suppose that it is desired to obtain an error relative to direct simulations in tissue of less than 5%, for both pressure and temperature. From the contours in Figs. 1(b) and 3(b), which pertain to a gain of 40, we search for values of N w /N t and e that will yield errors of less than 5% for both pressure and temperature. From Fig. 1(b), it can be see that a pressure error of less than 5% can only be attained with an e value less than about 0.5 and a value of N w /N t less than about 0.9. The choice (N w /N t, e) ¼ (0.7, 0.47) resides roughly at the center of the 5% error bands in both Figs. 1(b) and 3(b), and we make this selection. If the transducer gain G is significantly different from the values (20, 40, and 60) considered in Figs. 1 and 3, interpolation (or extrapolation) of the selected values of N w /N t and e can be performed, or the computations described in Sec. II can be performed using the exact gain desired. (3) Experiments in water are performed with the source acoustic pressure set so that N w ¼ N t (N w /N t ). In the present example, N w ¼ 0.3 0.7 ¼ 0.21. Pressure traces are recorded at axial locations spanning about the last 20% or 30% of the distance between the transducer and the 3440 J. Acoust. Soc. Am., Vol. 134, No. 5, November 2013 Dibaji et al.: Nonlinear derating with Gaussian modes

FIG. 3. (Color online) (a) Contours of relative difference between focal temperature rise predicted by the derating technique and by direct simulation in tissue, as a function of the parameters N w /N t and e. G ¼ 20. Coefficient of Nonlinearity in tissue N t ¼ 0.3. (b) Contours of relative difference between focal temperature rise predicted by the derating technique and by direct simulation in tissue, as a function of the parameters N w /N t and e. G ¼ 40. Coefficient of Nonlinearity in tissue N t ¼ 0.3. (c) Contours of relative difference between focal temperature rise predicted by the derating technique and by direct simulation in tissue, as a function of the parameters N w /N t and e. G ¼ 60. Coefficient of Nonlinearity in tissue N t ¼ 0.3. focus. For the present example, simulated pressures in water at the axial locations z ¼ 6.26 cm and 6.0 cm, corresponding to the focus and a slightly pre-focal location, are shown in Figs. 7(a) and 7(b). (4) In the focal plane, the acoustic pressure trace in water is also recorded at a radial location approximately 0.4 mm off axis. The off-axis pressure for the present location is shown in Fig. 8. J. Acoust. Soc. Am., Vol. 134, No. 5, November 2013 Dibaji et al.: Nonlinear derating with Gaussian modes 3441

FIG. 4. (Color online) Temperature rise as a function of exposure time determined by derating method with N w /N t ¼ 0.61 and e ¼ 0.38, along with temperature rise determined by direct simulation in tissue. G ¼ 40. Coefficient of Nonlinearity in tissue N t ¼ 0.3. FIG. 6. (Color online) (a). Pressure waveform at focus predicted by derating technique with reduced data set (0.8 < f < 1.0), compared with acoustic pressure waveform produced by direct simulation in tissue. N w /N t ¼ 0.61 and e ¼ 0.38. Coefficient of Nonlinearity in tissue N t ¼ 0.3. (b) Temperature rise at focus predicted by derating technique with reduced data set (0.8 < f < 1.0), compared with temperature rise produced by direct simulation in tissue. N w /N t ¼ 0.61 and e ¼ 0.38. Coefficient of Nonlinearity in tissue N t ¼ 0.3. FIG. 5. (a) Relative difference between focal acoustic pressure predicted by the derating technique and by direct simulation in tissue, as a function of e, for transducer gains of 20, 40, and 60. N w /N t ¼ 0.7. Nonlinearity in tissue N t ¼ 0.3. (b) Relative difference between focal temperature predicted by the derating technique and by direct simulation in tissue, as a function of e, for transducer gains of 20, 40, and 60. N w /N t ¼ 0.7. Nonlinearity in tissue N t ¼ 0.3. (5) Fourier transforms of the acoustic pressure vs time data are performed to obtain the source-derated modal amplitudes in water. Both real and imaginary parts are retained. The Fourier transform (real and imaginary parts) of the pressure measured at axial locations z ¼ 6.26 cm and 6.0 cm are shown in Figs. 9(a) and 9(b). It can be seen that on the order of 10 harmonics are significant at the focus, and roughly half that many at the pre-focal locations. We emphasize that these quantities represent only source-derated water amplitudes, and not represent the amplitude estimates w n, because they have not yet been endpoint derated. (6) The water modal amplitudes are then endpoint derated by multiplying by exp( neaz). The resulting amplitudes are the fully derated water amplitudes, w n. The w n values for the present example at the focal location are provided in Fig. 10. [The w n distribution at the pre-focal position (not shown) is similar to that in Fig. 10, though with less energy in the higher (n > 5) harmonics.] Endpoint derating has reduced the amplitudes to the point where only about half as many harmonics are now significant compared to the case of source-derating only [Fig. 9(a)]. 3442 J. Acoust. Soc. Am., Vol. 134, No. 5, November 2013 Dibaji et al.: Nonlinear derating with Gaussian modes

FIG. 7. (a) Focal pressure waveform in water for the example case of a 1.98 MHz transducer with a 5.0 cm diameter and a 6.26 cm focal length. Source pressure in water ¼ 0.25 MPa, source derated from a focal pressure in tissue of 0.34 MPa. (b) Same as in (a), but pressure plotted at the prefocal location z ¼ 6.0 cm. (7) The derated amplitudes w n are used to generate the modal products and sums in Eq. (8), resulting in an estimate for q n (f). (8) The source term q(f) is used in the numerical integral of Eq. (10) to obtain the final estimate for the tissue modal FIG. 8. Off-axis pressure (radial coordinate ¼ 0.04 cm) in the focal plane, for the example case of Fig. 7. FIG. 9. (a) Real and imaginary parts of the Fourier transform of the focal pressure signal in Fig. 7(a). These amplitudes represent water modes that are source-derated only. (b) Real and imaginary parts of the Fourier transform of the pre-focal pressure signal in Fig. 7(b). These amplitudes represent water modes that are source-derated only. amplitudes, a n. The a n values for the present example are provided in Fig. 11. [z ¼ 6.26 cm. Pre-focal values (not shown) are similar but with little energy above harmonic 4.] Relative to the derated water amplitudes w n (Fig. 10), there has been a slight shift in energy toward the higher harmonics. For example, comparing w n to a n at the focus, the fundamental mode has decreased about 30% while the third harmonic has increased more than 20% (Figs. 10 and 11). The predictions for a n for the current example are compared with values derived through direct simulation in tissue in Fig. 12. (9) The modal amplitudes are used in (7) to compute the tissue acoustic pressure, and (6) to determine temperature rise. The off-axis measurements in water would be used in (6b) to determine the beam widths required for the temperature computation. For the present example, the focal pressure and temperature rise predicted by the derating technique agree with direct simulations in tissue within 10%, in line with the agreement between the modal amplitudes in Fig. 12. We remind the reader that the data plotted in the contours all apply to the case N t ¼ 0.3. Contours for other N t J. Acoust. Soc. Am., Vol. 134, No. 5, November 2013 Dibaji et al.: Nonlinear derating with Gaussian modes 3443

FIG. 10. Water modes, w n, in the focal plane, for the example case of Fig. 7(a). These modal amplitudes represent water modes that are fully (sourceand endpoint-) derated. FIG. 12. Comparison of the modal amplitudes in tissue, a n, derived from the derating technique, with values computed from direct simulation in tissue, for the example case of Fig. 7(a). Axial location is the focal plane. values can be obtained from the authors. Alternatively, the interested reader can perform the computations necessary to solve Eqs. (5) (10), for the N t value of interest (and, as noted above, the exact G value of interest.) The computations involving different (N w /N t, e) combinations can be terminated when an acceptable level of accuracy (relative to a tissue simulation for the prescribed N t, G values) is obtained. The computation time required to simulate propagation in water, use the water modes to solve Eqs. (8) (10), and compare with the reference field comprised of direct simulation in tissue, is about 2 h on a 4-core personal computer with 24 GB memory. As simulations using different (N w /N t, e) combinations are performed, the comparison simulation in tissue need not be redone. In step 3 above, the amount of data acquisition necessary is dependent upon the transducer gain. The results of Fig. 6 show that an error of less than 10% in both acoustic pressure and temperature (relative to direct simulation in tissue) is possible when acoustic pressure measurements are made only in the last 20% of the axial locations (0.8 < f < 1), when the transducer gain was 40. Smaller FIG. 11. Estimates, a n, of the focal pressure-mode amplitudes in tissue, derived from the derating process, for the example case of Fig. 7(a). gains require larger data sets; approximately the last 25% is recommended for G ¼ 20. Reduced data sets for very low gains (G < 10) were not explored; measurement over the entire axial range can be performed to ensure accurate results. For gains larger than 40, data sets smaller than the last 20% may suffice, though data sets less than 20% of the full amount were not examined by the authors. From a computational standpoint, implementation of the method is straightforward, and requires little knowledge of numerical analysis by the user. Step 7 above involves algebraic manipulations of the derated modal amplitudes. Step 8 requires numerical quadrature, but the integrand is well behaved and simple methods suffice. As noted above, the trapezoidal rule was used in our computations. The accuracy of the derating method was comparable at all gain values studied, in the sense that (N w /N t, e) combinations can be found at all gain values where good accuracy (<10% difference with tissue simulations) was achievable. As an example in the low-gain regime, for a set of computations performed with G ¼ 10 and (N w /N t, e) ¼ (0.79, 0.07), the acoustic pressure and temperature errors were both determined to be 7%. In the sense that the space of (N w /N t, e) values for which the error is less than 10% increases with increasing gain, the method is more accurate at higher gain values. Besides applicability at low gains, an advantage of using the present approach to nonlinear derating is the direct availability of temperature estimates once the derated acoustic pressure amplitudes are determined. The modal beam widths are calculated from the off-axis acoustic pressure measurements and Eq. (6b), at which point Eq. (6a) provides the focal temperature as a function of time. For N t values on the order of 0.3, less than 10 modes typically contribute to the temperature sum. The Gaussian-mode approach becomes less accurate as the propagation becomes strongly nonlinear (Soneson and Myers 2007). The KZK approach, upon which the contours of Figs. 1 and 3 are based, also degrades in accuracy as the acoustic waveform becomes shocked, unless the model is 3444 J. Acoust. Soc. Am., Vol. 134, No. 5, November 2013 Dibaji et al.: Nonlinear derating with Gaussian modes

modified to treat shock formation and dissipation (Curra et al. 2000). Hence, the formulation presented in this paper is limited to moderately nonlinear procedures; we recommend an upper limit for the nonlinearity parameter in tissue of N t ¼ 0.3. Using typical transducer characteristics and tissue properties, this value corresponds to a power level of about 180 W when the gain is 40. Full validation of the derating technique presented in this paper requires, in addition to the above comparisons with numerical simulations in tissue, comparisons with experiments performed in tissue or a tissue phantom. These experiments are underway. Experimental challenges include selecting the proper value of r such that a sufficient number of radial modes can be resolved to make accurate temperature calculations. Accuracy (at the expense of increased measurement time) can be enhanced by making measurements at multiple radial locations, and curve fitting higherorder modes using multiple data points. Averaging over the width of the hydrophone may also prove problematic for higher harmonics. This issue is minimized when propagation is limited to the moderately nonlinear propagation regime identified above, where less than 10 harmonics contribute. V. CONCLUSION A nonlinear-derating technique based upon Gaussianmode theory proved to be a promising approach for characterizing HIFU systems. The technique was able to convert simulated acoustic pressure measurements in water to acoustic pressures and temperatures in tissue, with less than a 10% error when compared to results generated by direct numerical simulations in tissue. No assumption of Gaussian shading is required in the final results. Advantages of the derating approach include applicability over a wide range of gains, ease of computation (a single numerical quadrature is required), and readily obtained temperature estimates from the water measurements. Bessanova, O. V., Khokhlova, V. A., Canney, M. S., Bailey, M. R., and Crum L. A. (2010). A derating method for therapeutic applications of high intensity focused ultrasound, Acoust. Phys. 56, 354 363. Center for Devices and Radiological Health (2008). 510(k) guidance document Information for manufacturers seeking marketing clearance of diagnostic ultrasound systems and transducers, Report, U.S. Food and Drug Administration, Silver Spring, MD. Curra, F. P., Mourad, P. D., Khokhlova, V. A., Cleveland, R. O., and Crum, L. A. (2000). Numerical simulations of heating patterns and tissue temperature response due to high-intensity focused ultrasound, IEEE Trans. Ultrason. Ferroelectron. Freq. Control 47, 1077 1089. Hamilton, M. F., and Morfey, C. L. (1998). Model equations, in Nonlinear Acoustics, edited by M. F. Hamilton and D. T. Blackstock (Academic, San Diego), Chap. 3, pp. 40 63. Myers, M. R., and Soneson, J. E. (2009). Temperature modes for nonlinear Gaussian beams, J. Acoust. Soc. Am. 126, 425 433. Nyborg, W. L. (1988). Solutions of the bio-heat transfer equation, Phys. Med. Biol. 33(7), 785 792. Nyborg, W. L., and Wu, J. (1994). Solution of the linear bio-heat transfer equation, Phys. Med. Biol. 39(5), 924 926. Rielly, M., Humphrey, V., and Duck, F. (2000). A theoretical and experimental investigation of nonlinear ultrasound propagation through tissue mimicking fluids, IEEE Ultrason. Symp. Proc. 2, 1355 1358. Schafer, M. E. (1990). Alternative approaches to in-situ intensity estimation, Proc. IEEE Ultrason. Symp. 3, 1381 1384. Soneson, J. E. (2008). A user-friendly software package for HIFU simulation, AIP Conf. Proc. 1113, 165 169. Soneson, J. E., and Myers, M. R. (2007). Gaussian representation of highintensity focused ultrasound beams, J. Acoust. Soc. Am. 122, 2526 2531. Szabo, T. L., Clougherty, F., and Grossman, C. (1999). Effects of nonlinearity on the estimation of in situ values of acoustic output parameters, J. Ultrasound Med. 18, 33 41. Verma, P. K., Humphrey, V. F., and Duck, F. A. (2005a). Broadband attenuation and nonlinear propagation in biological fluids: An experimental facility and measurements, Ultrasound Med. Biol. 31, 1723 1733. Verma, P. K., Humphrey, V. F., and Duck, F. A. (2005b). Broadband measurements of the frequency dependence of attenuation coefficient and velocity in amniotic fluid, urine, and human serum albumin solutions, Ultrasound Med. Biol. 31, 1375 1381. Wu, J., and Du, G. (1990). Temperature elevation generated by a focused Gaussian beam of ultrasound, Ultrasound Med. Biol. 16(5), 489 498. Wu, J., and Nyborg, W. L. (1992). Temperature rise generated by a focussed Gaussian beam in a two-layer medium, Ultrasound Med. Biol. 18(3), 293 302. J. Acoust. Soc. Am., Vol. 134, No. 5, November 2013 Dibaji et al.: Nonlinear derating with Gaussian modes 3445