Open Access
29 June 2016 Microwave-heating-coupled photoacoustic radar for tissue diagnostic imaging
Author Affiliations +
Abstract
An investigation of microwave (MW) heating effects on biotissue for enhancing photoacoustic radar (PAR) signals was conducted. Localized tissue heating generated by MWs was used to improve PAR imaging depth and signal-to-noise ratio (SNR). Elevated temperatures were measured with thermocouples in ex vivo bovine muscle. The measured temperature rise on the heated spot surface by MWs was in agreement with theoretical predictions. The study showed localized MW heating can increase the photoacoustic imaging depth by 11%, and the SNR by 5% in ex vivo bovine muscle.

1.

Introduction

Photoacoustic (PA) imaging combines the advantages of high contrast of light absorption, high resolution, and penetration depth of ultrasound, and has been investigated for breast cancer detection in recent years.1,2 The performance characteristics of PA imaging systems include the maximum imaging depth, sensitivity, resolution, and contrast.3,4 PA imaging has good sensitivity, resolution, and contrast, but the major limitation is the optical penetration depth.35 Light intensity decreases exponentially with penetration depth because of strong scattering and absorption in biological tissues.6,7 The optical window in the range of 700 to 900 nm allows relatively higher light penetration depth.6,7 PA signals can be improved by contrast agents, such as silica-coated super paramagnetic iron oxide nanoparticles,8 system optimization, signal processing, or higher laser power.9 However, contrast agents have side effects, system optimization increases the complexity and cost, and higher laser power exposure induces potential damage to skin.

Manipulation of tissue temperature has been used for thermal ablation, hyperthermia, and imaging.1013 The normal body temperature of a healthy adult is 37°C.14,15 Studies show that most normal healthy tissues such as liver, kidney, and muscles can withstand temperatures up to 44°C for 30 min.10,1618 The relationship between thermal exposure time and temperature in most normal healthy tissues has also been studied.18,19 The reported results show that heating effects of 60-min exposure time at 43°C are equivalent to 30 (15) min of exposure at 44°C (45°C). These and other results indicate that there exist ranges of maximum safe exposure times at substantially higher temperatures than normal for most normal healthy tissues,16,17,20 and thus the possibility of using high temperature for significant improvements in the PA signal strength and imaging depth. Thermal energy can be locally delivered by microwaves (MWs) in biological tissues.11 An MW-energy applicator (antenna) is placed on the target site to increase the temperature surrounding the applicator inside the tissue. Coaxial slot antennas are the most popular tools for MW heating purposes due to their small dimensions and low cost.21,22

Generally speaking, two kinds of laser sources are used to generate PA signals: (1) pulsed and (2) continuous wave (CW)-modulated laser sources. The pulsed laser is the conventional laser source, but CW lasers have the advantages of low cost, portability, and narrow band detection [i.e., high signal-to-noise ratio (SNR)].9 The photoacoustic radar (PAR) uses chirp signals for laser-beam modulation and matched filtering for signal processing.23 Temperature can affect PAR signals, which become stronger at higher temperatures.12,13 The imaging depth and SNR can be improved through uniform heating.12,13 Most research reports of MW heating have been in applications of surgical ablation.21,22,2427 Our earlier study12 showed that “under uniform heating,” the thermally enhanced PA radar probe can increase the signal strength from ex vivo beef muscle by 13% and the imaging depth over 20% with a temperature increase from 37°C to 43°C. However, imaging improvement through uniform heating takes very long and can cause discomfort and inconvenience to patients. Thus, it is necessary to find a localized heating method for precise tissue targeting and shortening the exposure time. The present feasibility study investigates the thermal effects on PAR signal enhancement “under targeted (localized)” heating of biotissues in the same temperature range using MW heating. The tradeoff was found to be rapid (localized MW) heating at relatively low signal strength increase versus very slow uniform heating with significantly higher signal strength increase.

2.

Theoretical Background

PAR signals were generated by harmonically modulating light illumination on absorbers.28 The response amplitude spectrum p˜(ω) of the generated acoustic pressure can be written in one dimension as12,23,29

Eq. (1)

|p˜(ω)|=μaΓeμefzμa2ca2+ω2E0,
where μa represents the optical absorption coefficient (m1), ω is the angular modulation frequency (rad/s), Γ is the Grüneisen parameter (dimensionless), μef is the effective optical attenuation coefficient (m1), ca is the speed of sound (m/s), z is the subsurface distance (m), and E0 is the optical fluence at the surface of the absorber (J·Hz/m2). The Grüneisen parameter is defined as

Eq. (2)

Γ=βca2cp,
where cp is the specific heat capacity [J/(kg·°C)]. The matched-filter pulse compression method was introduced into the PAR system to give the frequency-domain PA signal necessary depth resolution and adequate SNR.28 Matched filtering is introduced by the cross-correlation method to measure the similarity of two signals (frequency content and phase). The cross-correlation output is

Eq. (3)

R(t)=+r(τt)·s(τ)dτ,
where s(t) and r(t) are the detected and reference signals, respectively. Linear frequency modulation (LFM) is the simplest and most popular signal pulse compression method, which can reduce side lobes and improve depth resolution, among other attributes.28,30 An LFM signal can be expressed as

Eq. (4)

rt=rect(t2K)cos(ω0t+πBKt2),K2tK2,
where ω0 is the center angular frequency, K is the pulse duration, and B is the frequency bandwidth. The output of a matched filter can be written as

Eq. (5)

Rt=K2rect(t2K)sin[πBt(1tK)]πBtcos(ω0t).
In frequency-domain signal processing, the Fourier transform of the reference signal rt is r˜(ω) and the complex conjugate of r˜(ω) is h˜(ω). The matched filter output R˜(ω) in the frequency domain is obtained as

Eq. (6)

R˜(ω)=s˜(ω)·h˜(ω).

The spectrum of the LFM signal rt is

Eq. (7)

r˜(ω)=12K2K2ej[(ω0ω)t+πBKt2]dt+12K2K2ej[(ω0ω)t+πBKt2]dt.

The Fresnel integrals C(X) and S(X) are defined by

Eq. (8)

C(X)=0Xcosπy22dy,

Eq. (9)

S(X)=0Xsinπy22dy.

It can be shown that

Eq. (10)

r˜(ω)=12K2B{ej(ωω0)24πB[C(x3)+C(x4)+jS(x3)+jS(x4)]}+12K2B{ej(ω+ω0)24πB[C(x5)+C(x6)jS(x5)jS(x6)]},
where

Eq. (11)

x3=Ωπ(K2+ωω0Ω),

Eq. (12)

x4=Ωπ(K2ωω0Ω),

Eq. (13)

x5=Ωπ(K2+ω+ω0Ω),

Eq. (14)

x6=Ωπ(K2ω+ω0Ω),
where Ω=(2πB/K). Thus, the complex conjugate of r˜(ω) is given as

Eq. (15)

h˜(ω)=12K2B{cos(ωω0)24πB[C(x3)+C(x4)]+cos(ω+ω0)24πB[C(x5)+C(x6)]+sin(ωω0)24πB[S(x3)+S(x4)]+sin(ω+ω0)24πB[S(x5)+S(x6)]}+j2K2B{sin(ωω0)24πB[C(x3)+C(x4)]sin(ω+ω0)24πB[C(x5)+C(x6)]cos(ωω0)24πB[S(x3)+S(x4)]+cos(ω+ω0)24πB[S(x5)+S(x6)]}.

The amplitude of h˜(ω) and s˜(ω) can be derived as

Eq. (16)

|h˜(ω)|=|s˜(ω)|12K2B[C(x3)+C(x4)]2+[C(x5)+C(x6)]2+[S(x3)+S(x4)]2+[S(x5)+S(x6)]2.

Fresnel integrals can be approximated as31

Eq. (17)

C(x)12+1πxsin(πx22),for  x1,

Eq. (18)

S(x)121πxcos(πx22),for  x1.

For frequencies higher than 300 kHz, the Fresnel integrals are 1/2. Therefore, the final matched filter output amplitude spectrum of two signals h˜(ω) and s˜(ω) is approximately32 given by

Eq. (19)

|s˜(ω)·h˜(ω)|K2B.

The output amplitude spectrum is proportional to the duration of the chirp and inversely proportional to the bandwidth of the chirp. For narrow frequency bandwidth B and long duration time K chirps, the amplitude increases.

Water plays an important role in biotissues. The total water content of human tissues can be up to 77%.33 The Grüneisen parameter of water increases monotonically from 0.11 at 20°C to 0.24 at 45°C [shown in Fig. 1(a)]. The thermal expansion coefficient increases 107% from 20°C to 45°C.34 The specific heat capacity of water is a mildly temperature-dependent parameter and the variation is less than 1% from 20°C to 45°C.35 The speed of sound in water increases with temperature from 1482 to 1536  m/s, which is 3.6% increase from 20°C to 45°C.36 Figure 1(b) shows a comparison of the temperature dependence of the parameters constituting the Grüneisen constant. It can be seen that the thermal expansion coefficient contributes much more than the other two parameters to the increase.

Fig. 1

(a) Grüneisen parameter of water dependence on temperature and (b) increase rate comparison of Grüneisen parameter elements of water from 20°C to 45°C.

JBO_21_6_066018_f001.png

Bovine muscles are used as samples for experimental validation. Myoglobin and hemoglobin are the main light absorbers in bovine muscle tissues and the concentration of myoglobin is much higher than hemoglobin:37 Myoglobin content is 0.41  mmol/kg, whereas hemoglobin content is only 0.066  mmol/kg in bovine muscle.37 Myoglobin concentration can determine the color of muscle: higher myoglobin concentration yields darker color.38 In ex vivo tissues, myoglobin and hemoglobin are present in oxygenated or deoxygenated forms, but deoxygenated concentrations are higher. On the surface of ex vivo tissues, the degree of oxygenated concentrations is higher. After myoglobin is oxygenated (oxymyoglobin), color changes from dark purple to bright red, with the muscle assuming a bright cherry-red color.38

A typical composition of bovine muscle consists approximately of 75% water, 19% protein, 3% fat, 1% fiber, 1% carbohydrate, and 1% ash.39,40 Temperature-dependent empirical thermal property models of individual biotissue-components have been introduced for the Grüneisen parameter estimation.12,41 The Grüneisen parameter of a typical bovine muscle dependence on temperature has been studied from 20°C to 45°C.12 In this temperature range, the increase is 67% for the Grüneisen parameter, 1.2% for the speed of sound, 0.33% for the specific heat capacity, and 64% for the thermal expansion coefficient.12 The highest increase is, again, the thermal expansion coefficient.

The PAR signal changes with the Grüneisen parameter, and thus becomes temperature dependent.4244 Temperature rise in tissue depends on both heating power and time, and is also related to tissue properties and heat transfer to the vasculature.12 The specific absorption rate (SAR) describes the time-averaged amount of MW power deposited per unit mass of tissue (W/kg). The SAR of electromagnetic energy incident on a tissue sample is defined by

Eq. (20)

SAR=σ2ρ|E|2,
where σ is the electrical conductivity of biological tissue (S/m), ρ is its density (kg/m3), and |E| is the amplitude of the applied electric field (V/m). The SAR determines the heating pattern of the MW applicator (antenna) in tissue. Its depth distribution can be approximately expressed as an exponential profile45

Eq. (21)

SAR=S0eμz,
where S0 represents the amplitude of SAR at the surface, z is the distance from the heated spot (m), and μ represents the MW energy absorption coefficient (m1).

The MW-generated heat is strongly dependent on the absorption coefficient of tissue, and on the frequency and intensity of the energy source.46 Heat can be reduced by increasing perfusion with temperature.47 Assuming that tissue is electromagnetically heterogeneous and acoustically homogeneous, the Pennes bioheat transfer equation was developed for roughly describing temperature elevation in biological tissues47

Eq. (22)

ρcpTt=k2Tρbcbwb(TTb)+q,
where T is the tissue temperature (°C), cp is the tissue specific heat capacity [J/(kg·°C)], ρ is the tissue density, k is the tissue thermal conductivity [W/(m·°C)], wb is the blood flow rate through the tissue [m3/(s·kg)], ρb is the blood density, cb is the blood specific heat capacity, Tb is the blood temperature (°C), t is the heating time (s), and q is the volume heating source (W/m3). All temperature information about blood perfusion is included in the term ρbwbcb(TTb). For ex vivo experiments, the blood perfusion term wbcb=0. Due to the efficiency of MW heating, the heating time is less than half a minute. So for in vivo study, the blood flow effects can be ignored in such short time for tissues having low blood flow velocity, such as skeletal muscles. Thus, Eq. (22) can be simplified as48

Eq. (23)

ρcpTt=k2T+q.

For MW heating, the geometry is shown in Fig. 2(a), where the cylindrical coordinate system is used. (r,z,t), (r,z,τ) are the observation and source coordinates, respectively. The translated temperature of the thermal field in the medium is written as

Eq. (24)

θ(r,z,t)T(r,z,t)T0,
where T0 is the ambient temperature. The translated thermal field satisfies the following equation:

Eq. (25)

ρcpθt=krr(rθr)+k2θz2+q(r,z).

Fig. 2

(a) Experimental setup for MW heating and (b) geometry of MW antenna.

JBO_21_6_066018_f002.png

The initial and boundary conditions for MW heating are given by

Eq. (26)

θ(r,z,0)=0,

Eq. (27)

θ(r,0,t)z=0,

Eq. (28)

q(z)=q0eμz,

Eq. (29)

q0=η·ρ·S0,
where q0 is the heat flux at the surface z=0 and η is the heat generation efficiency. Here q is the MW source power, taken to be only a function of the depth (axial) coordinate and assumed uniform along the radial dimension, consistent with our experiments. The solution of the boundary value problem Eqs. (25) to (29) can be derived using the Green function (GF) method

Eq. (30)

θ(r,z,t)=1ρcpτ=0tr=0bz=0Lq(z)G(r,z,t|r,z,τ)2πrdτdrdz.

For the geometry shown in Fig. 2(a), it is assumed that the absorbing medium is cylindrical and semi-infinite lengthwise, 0<r<b in the radial direction, and the penetration depth is L=1/μ in the axial direction, corresponding to the absorbing area of the MW energy. The GF is separated into the product of a radial and an axial function49

Eq. (31)

G(r,z,t|r,z,τ)=GR(r,t|r,τ)GA(z,t|z,τ).

The radial function GR satisfies

Eq. (32)

GR(r,t|b,0)=GR(r,t|0,τ)=0,
where GR is given by50

Eq. (33)

GR(r,t|r,τ)=14πα(tτ)exp[(r2+r2)4α(tτ)]×I0[rr2α(tτ)],
where I0 is the modified Bessel function of the first kind of order zero and α is the thermal diffusivity (α=(k/ρcp)). The axial function satisfies the following condition:

Eq. (34)

zGA(z,t|z,τ)|z=0=0,
and is given by50

Eq. (35)

GA(z,t|z,τ)=14πα(tτ){exp[(zz)24α(tτ)]+exp[(z+z)24α(tτ)]}.
Thus, the solution of Eq. (30) is

Eq. (36)

θ(r,z,t)=1ρcpτ=0tdτr=0bGR(r,t|r,τ)2πrdrz=0Lq(z)GA(z,t|z,τ)dz.
This yields

Eq. (37)

θ(r,z,t)=q0ρcpτ=0tdτr=0b2πr4πα(tτ)exp[(r2+r2)4α(tτ)]×I0[rr2α(tτ)]dr×z=0L14πα(tτ){exp[μz+(zz)24α(tτ)]+exp[μz+(z+z)24α(tτ)]}dz.
For small values of b, at r=0, z=0, the solution is simplified as50

Eq. (38)

θ(r,z,t)|r=0z=0=q0kπ(αt)1/2.

Tumors contain higher water content than normal healthy tissues,51 and thus they have higher electrical conductivity than low-water-content healthy tissue.5254 Studies have shown that their electrical conductivity is about 10% higher than normal tissue.55,56 This suggests that MWs can raise the temperature of tumor tissues higher than that of the host mammary tissues, and thus generate higher contrast between tumors and normal tissues. The MW penetration depth is the inverse of the microwave absorption coefficient, where the electric field power is reduced by 1/e246 at depths between 8.5 cm for bovine fat tissue and 1.1 cm for bovine muscle tissue at 2.45 GHz, respectively.46,52,57 Penetration depths in other tissues have values between those of fat and muscle.

3.

Experimental Setups

The experimental setup for MW heating used in this work consisted of a MW generator (SSG-4000HP, Mini-circuit), a 40-dB power amplifier (ZHL-16W-43+, Mini-circuit), a coaxial slot antenna, thermocouples, and a data-acquisition device (USB-6221, NI) as shown in Fig. 2(a). The generated MW signals were amplified with a power amplifier and then sent to the coaxial slot antenna. Since water is a good absorber of MW energy, an external waveguide applicator has difficulty in delivering the MW energy to the absorber immersed in the water (our setup), so a single-slot coaxial antenna was employed for this research. This type of antenna can locally produce an electromagnetic field around the slots and heat the target area (deep inside the absorber) efficiently.21,22,24 The electrical power supplied to the antenna was set at 2 W. The frequency was set at 2.45 GHz, the most popular frequency used for hyperthermia and heating applications.21,22,24 The coaxial slot antenna was constructed from a RG402 semirigid coaxial copper cable as shown in Fig. 2(b). The outer conductor and core were both made of copper. A 2-mm ring-shaped slot was cut by removing a part of the outer conductor near the tip of the cable. The outer and inner conductors were connected at the tip by soldering. The reflected signal was sampled at the reflected port of a directional coupler, and measured by a network analyzer (E5063A, Agilent). The measured power loss was about 12  dB (about 5% loss). Fine-wire thermocouples (K-type, Omega) were employed for temperature measurements. The measured data were collected by a USB DAQ device (USB-6221, NI), and then sent to a computer synchronized with the MW generator.

The experimental setup diagram for the MW-enhanced PAR system is shown in Fig. 3(a). A CW diode laser emitting at 800 nm was employed. The light beam size was 3.5 mm and was modulated by linear frequency waveform chirps (0.3 to 1.3 MHz, 1 ms in duration) generated by the signal-generation card NI PXI-5421 (National Instruments, Austin, Texas). A focused ultrasound transducer (Panametrics-NDT, V314) was utilized as a detector. The detected signals were amplified with a preamplifier (Panametrics-NDT, 5676) first, and then were sent to the digital data-acquisition card NI PXIe-5122 (National Instruments, Austin, Texas) in a computer. The sampling rate of the data-acquisition card NI PXIe-5122 was 100  MS/s, and the maximum input voltage was 10 V. Bovine muscle was purchased from a local store and was stored in a refrigerator at 4°C. Most experiments were performed within 3 days from the purchase. Fresh samples used for each experiment were cut from the same part. The sample dimensions were 30-mm long, 30-mm wide, and 20-mm thick. The samples were all wrapped in thin transparent plastic wrap during the experiments.

Fig. 3

(a) Experimental setup of the MW coupled PAR system and (b) measured temperature on a heated surface spot of bovine muscle versus time compared with the analytically estimated MW heating temperature.

JBO_21_6_066018_f003.png

4.

Results and Discussion

Figure 3(b) shows the comparison of the experimentally measured temperature in ex vivo bovine muscle of the MW-heated surface spots with the estimated temperature using Eq. (38). For bovine muscle at 37°C, the density is 1073.445  kg/m3,12,41 the specific heat capacity is 3632.7  J/(kg·°C),12,41 and the thermal conductivity is 0.5228  W/(m·°C).41,46 These parameters are all temperature dependent, but the variations are all less than 2% from 37°C to 45°C,12,41 and are thus ignored in this study. The thermal diffusivity of bovine muscle can be calculated from the defining equation α=(k/ρcp). This yields α=1.34×107  m2/s in the range 37°C to 45°C, so variation with T is ignored. For the 30-s heating period in Fig. 3(b), all the absorbed energy was assumed converted to heat (η=100%), and the experimentally measured temperature at the heated surface spot was in excellent agreement with the analytically estimated temperature. The SAR was estimated to be 3.1  W/kg (the best-fit value of the theory to the experimental results). The estimated error between measured and theoretical results was less than 10%. The temperature rose to 44.5°C from 37°C after 30-s of MW heating, a 20% increase. No visible damage was observed after the heating experiment.

Figure 4(a) shows the amplitude of the cross-correlation peak between the reference (by a uniform heating method) and the measured PAR signals under the influence of MW heating. The results showed that the PAR signal strength increased by about 8% with MW heating at ambient temperature 37°C and the SNR increased by 5%. The average elevated temperature during the PAR measurement was 42°C. In comparison, with uniform heating, the PAR-signal-amplitude increased up to 13% from 37°C to 42°C with the sample immersed in heated water. For MW heating-assisted PAR, the nonuniformly distributed temperature field in tissue was generated by the MW applicator, and at lower temperatures, the PA signals were weaker.12 Therefore, the nonuniform temperature distribution and the lower heating efficiency under MW heating were the reasons for the less-efficient PAR signal increase. Theoretically, the SNR can be improved by a further 8% with a more efficient thermally assisted method,12,13 the single most important criterion for such efficiency being achieving temperature uniformity. The experimental results of the imaging depth enhancement study are shown in Fig. 4(b). The sample was placed in a box filled with 0.47% intralipid solution simulating an optical scatterer as shown in Fig. 4(c).12,58 The estimated scattering coefficient of the 0.47% intralipid solution was 0.014  mL1Lmm1 (concentration×sample thickness) and the absorption coefficient was 0.003  mL1Lmm1.12,58 The box was translated with microstages to attain different subsurface absorber depths. At 18 mm, the PAR signal was measured with large errors (20%), and the noise was high. With MW heating, the PAR signal became stronger (about 8% increase). The experimental results show that the imaging depth was increased by 11% with MW heating assistance.

Fig. 4

(a) Measured PAR cross-correlation amplitude peak from ex vivo bovine muscle with MW and uniform heating, (b) PAR signal imaging depth study under MW heating, and (c) experimental setup for PAR imaging depth study (the absorber surrounding medium was 0.47% intralipid solution).

JBO_21_6_066018_f004.png

Figure 5 shows an experimentally one-line-scanned image of a 4-mm wide ex vivo bovine muscle sample. Figure 5(a) is the picture of the sample as it was placed between clamps. Figure 5(b) shows a bright PAR horizontal-line image at ambient temperature 37°C with intralipid solution as the scatterer. The sample was placed 18 mm below the intralipid surface. The scanning signals were collected while moving the laser beam and transducer in tandem across the sample. Figure 5(c) shows the respective image generated with additional MW heating. Comparing Figs. 5(b) and 5(c), the MW-assisted image in Fig. 5(c) appears brighter than in Fig. 5(b) with enhanced contrast against a receding background due to the stronger signal (8% increase).

Fig. 5

(a) Picture of a 4-mm wide ex vivo bovine muscle sample, (b) one-dimensional image at 37°C with optical scattering, and (c) same imaging conditions as in (b) with added MW heating.

JBO_21_6_066018_f005.png

5.

Conclusions

A feasibility investigation of localized MW heating effects on PAR signal enhancement in ex vivo bovine muscle was conducted. The localized MW heating was experimentally and theoretically investigated. The calculated and measured results showed agreement at the heated spot over the same heating period and they demonstrated that MW sources can generate a thermal field at low-input-power in bovine muscle for enhancing PAR signals. The PAR SNR increased by about 5% with MW heating assistance for tissue temperature rise up to 42°C, whereas the maximum imaging depth was improved by 11%. MW heating was thus shown to be an effective heating source for biotissues. In view of earlier tissue heating studies,12 it is concluded that MW heating can be used as a substantially more significant depth resolution enhancement method for PAR imaging of tissues, resulting in up to 13% increase in signal strength and over 20% increase in imaging depth under a localized temperature increase from 37°C to 43°C, provided the MW source can be optimized to generate more uniform subsurface tissue heating around the investigated tissue location than the system developed for this study.

Acknowledgments

A. M. gratefully acknowledges the support of NSERC through a Discovery Grant and the support of the Samsung Advanced Institute of Technology (SAIT) through a Global Research Outreach (GRO) Research Project Grant.

References

1. 

R. O. Esenaliev et al., “Laser optoacoustic imaging for breast cancer diagnostics: limit of detection and comparison with X-ray and ultrasound imaging,” Proc. SPIE, 2979 71 –82 (1997). http://dx.doi.org/10.1117/12.280213 PSISDG 0277-786X Google Scholar

2. 

T. Khamapirad et al., “Diagnostic imaging of breast cancer with LOIS: clinical feasibility,” Proc. SPIE, 5697 35 –44 (2005). http://dx.doi.org/10.1117/12.596916 PSISDG 0277-786X Google Scholar

3. 

X. Wang et al., “Noninvasive laser-induced photoacoustic tomography for structural and functional in vivo imaging of the brain,” Nat. Biotechnol., 21 803 –806 (2003). http://dx.doi.org/10.1038/nbt839 NABIF9 1087-0156 Google Scholar

4. 

V. G. Andreev, A. A. Oraevsky and A. A. Karabutov, “Wide-band acoustic pulse detection in opto-acoustic tomography system,” in IEEE Ultrasonics Symp. 2000, 1205 –1208 (2000). http://dx.doi.org/10.1109/ULTSYM.2000.921539 Google Scholar

5. 

M. Xu and L.V. Wang, “Photoacoustic imaging in biomedicine,” Rev. Sci. Instrum., 77 041101 (2006). http://dx.doi.org/10.1063/1.2195024 RSINAK 0034-6748 Google Scholar

6. 

W. F. Cheong, S. A. Prahl and A. J. Welch, “A review of the optical properties of biological tissues,” IEEE J. Quantum Electron., 26 2166 –2185 (1990). http://dx.doi.org/10.1109/3.64354 Google Scholar

7. 

N. Won et al., “Imaging depths of near-infrared quantum dots in first and second optical windows,” Mol. Imaging, 11 (4), 338 –352 (2012). http://dx.doi.org/10.2310/7290.2011.00057 Google Scholar

8. 

R. Alwi et al., “Silica-coated super paramagnetic iron oxide nanoparticles (SPION) as biocompatible contrast agent in biomedical photoacoustics,” Biomed. Opt. Express, 3 2500 –2509 (2012). http://dx.doi.org/10.1364/BOE.3.002500 BOEICL 2156-7085 Google Scholar

9. 

B. Lashkari and A. Mandelis, “Comparison between pulsed laser and frequency-domain photoacoustic modalities: signal-to-noise ratio, contrast, resolution, and maximum depth detectivity,” Rev. Sci. Instrum., 82 094903 (2011). http://dx.doi.org/10.1063/1.3632117 RSINAK 0034-6748 Google Scholar

10. 

H. I. Robins et al., “A nontoxic system for 41.8 degrees C whole-body hyperthermia: results of a Phase I study using a radiant heat device,” Cancer Res., 45 (8), 3937 –3944 (1985). Google Scholar

11. 

J. Stang et al., “A preclinical system prototype for focused microwave thermal therapy of the breast,” IEEE Trans. Biomed. Eng., 59 2431 –2438 (2012). http://dx.doi.org/10.1109/TBME.2012.2199492 IEBEAX 0018-9294 Google Scholar

12. 

W. Wang and A. Mandelis, “Thermally enhanced signal strength and SNR improvement of photoacoustic radar module,” Biomed. Opt. Express, 5 2785 –2790 (2014). http://dx.doi.org/10.1364/BOE.5.002785 BOEICL 2156-7085 Google Scholar

13. 

W. Wang and A. Mandelis, “Thermally enhanced photoacoustic radar imaging of biotissues,” Int. J. Thermophys., 36 900 –904 (2015). http://dx.doi.org/10.1007/s10765-015-1838-y IJTHDY 0195-928X Google Scholar

14. 

G. Kelly, “Body temperature variability (Part 1): a review of the history of body temperature and its variability due to site selection, biological rhythms, fitness, and aging,” Altern. Med. Rev., 11 (4), 278 –293 (2006). Google Scholar

15. 

P. A. Mackowiak and G. Worden, “Carl Reinhold August Wunderlich and the evolution of clinical thermometry,” Clin. Infect. Dis., 18 458 –467 (1994). http://dx.doi.org/10.1093/clinids/18.3.458 CIDIEL 1058-4838 Google Scholar

16. 

L. F. Fajardo, “Pathological effects of hyperthermia in normal tissues,” Cancer Res., 44 (10 Suppl), 4826s –4835s (1984). Google Scholar

17. 

J. van der Zee, “Heating the patient: a promising approach?,” Ann. Oncol., 13 (8), 1173 –1184 (2002). http://dx.doi.org/10.1093/annonc/mdf280 ANONE2 0923-7534 Google Scholar

18. 

S. A. Sapareto and W. C. Dewey, “Thermal dose determination in cancer therapy,” Int. J. Radiat. Oncol. Biol. Phys., 10 787 –800 (1984). http://dx.doi.org/10.1016/0360-3016(84)90379-1 Google Scholar

19. 

M. R. Horsman and J. Overgaard, “Hyperthermia: a potent enhancer of radiotherapy,” Clin. Oncol., 19 418 –426 (2007). http://dx.doi.org/10.1016/j.clon.2007.03.015 Google Scholar

20. 

P. Sminia et al., “Effect of hyperthermia on the central nervous system: a review,” Int. J. Hyperthermia, 10 1 –30 (1994). http://dx.doi.org/10.3109/02656739409009328 Google Scholar

21. 

C. L. Brace, “Dual-slot antennas for microwave tissue heating: parametric design analysis and experimental validation,” Med. Phys., 38 4232 –4240 (2011). http://dx.doi.org/10.1118/1.3601019 MPHYA6 0094-2405 Google Scholar

22. 

S. Kikuchi et al., “Control of heating pattern for interstitial microwave hyperthermia by a coaxial-dipole antenna-aiming at treatment of brain tumor,” Electron. Commun. Jpn., 90 31 –38 (2007). http://dx.doi.org/10.1002/(ISSN)1520-6424 Google Scholar

23. 

B. Lashkari and A. Mandelis, “Linear frequency modulation photoacoustic radar: optimal bandwidth and signal-to-noise ratio for frequency-domain imaging of turbid media,” J. Acoust. Soc. Am., 130 1313 –1324 (2011). http://dx.doi.org/10.1121/1.3605290 JASMAN 0001-4966 Google Scholar

24. 

P. Keangin, P. Rattanadecho and T. Wessapan, “An analysis of heat transfer in liver tissue during microwave ablation using single and double slot antenna,” Int. Commun. Heat Mass Transfer, 38 757 –766 (2011). http://dx.doi.org/10.1016/j.icheatmasstransfer.2011.03.027 Google Scholar

25. 

M. G. Skinner et al., “A theoretical comparison of energy sources-microwave, ultrasound and laser-for interstitial thermal therapy,” Phys. Med. Biol., 43 3535 –3547 (1998). http://dx.doi.org/10.1088/0031-9155/43/12/011 PHMBA7 0031-9155 Google Scholar

26. 

W. Chen et al., “Differences in the lesion formation process between focused ultrasound and microwave ablations,” Med. Phys., 33 1346 –1351 (2006). http://dx.doi.org/10.1118/1.2193689 MPHYA6 0094-2405 Google Scholar

27. 

V. Frenkel, “Ultrasound mediated delivery of drugs and genes to solid tumors,” Adv. Drug Delivery Rev., 60 1193 –1208 (2008). http://dx.doi.org/10.1016/j.addr.2008.03.007 ADDREP 0169-409X Google Scholar

28. 

S. A. Telenkov and A. Mandelis, “Fourier-domain biophotoacoustic subsurface depth selective amplitude and phase imaging of turbid phantoms and biological tissue,” J. Biomed. Opt., 11 044006 (2006). http://dx.doi.org/10.1117/1.2337290 JBOPFO 1083-3668 Google Scholar

29. 

V. Ė Gusev and A. A. Karabutov, Laser Optoacoustics, American Institute of Physics, New York (1993). Google Scholar

30. 

C. E. Cook, Radar Signals: An Introduction to Theory and Application, Artech House, Boston, Massachusetts (1993). Google Scholar

31. 

B. R. Mahafza, Radar Systems Analysis and Design using MATLAB, 2nd ed.Chapman & Hall/CRC, Boca Raton, Florida (2005). Google Scholar

32. 

W. Wang, “Signal strength and SNR enhancement techniques for frequency domain photoacoustic radar imaging,” Univ. of Toronto, (2015). Google Scholar

33. 

V. A. Dubinskaya et al., “Comparative study of the state of water in various human tissues,” Bull. Exp. Biol. Med., 144 294 –297 (2007). http://dx.doi.org/10.1007/s10517-007-0314-5 BEXBAN 0007-4888 Google Scholar

34. 

N. A. Lange, Lange’s Handbook of Chemistry, 11th ed.McGraw-Hill, New York (1973). Google Scholar

35. 

N. Osborne, H. Stimson and D. Ginnings, “Measurements of heat capacity and heat of vaporization of water in the range 0 degrees to 100 degrees C,” J. Res. Nat. Bur. Stand., 23 197 –260 (1939). http://dx.doi.org/10.6028/jres.023.008 Google Scholar

36. 

S. Temkin, Elements of Acoustics, John Wiley, New York, Toronto (1981). Google Scholar

37. 

J. R. Zijp and J. J. ten Bosch, “Optical properties of bovine muscle tissue in vitro; a comparison of methods,” Phys. Med. Biol., 43 3065 –3081 (1998). http://dx.doi.org/10.1088/0031-9155/43/10/026 PHMBA7 0031-9155 Google Scholar

38. 

L. M. L. Nollet et al., Handbook of Meat, Poultry and Seafood Quality, 1st ed.Blackwell Publishing, Ames, Iowa (2007). Google Scholar

39. 

R. A. Lawrie, Lawrie’s Meat Science, 7th ed.CRC Press, Boca Raton, Florida (2006). Google Scholar

40. 

A. M. Pearson and R. B. Young, Muscle and Meat Biochemistry, Academic Press, San Diego, California (1989). Google Scholar

41. 

Y. Choi, M. R. Okos, “Effects of temperature and composition on the thermal properties of foods,” Food Engineering and Process Applications: Transport Phenomena, 1 93 –101 Elsevier Applied Science, Amsterdam, The Netherlands (1986). Google Scholar

42. 

S. M. Nikitin, T. D. Khokhlova and I. M. Pelivanov, “Temperature dependence of the optoacoustic transformation efficiency in ex vivo tissues for application in monitoring thermal therapies,” J. Biomed. Opt., 17 061214 (2012). http://dx.doi.org/10.1117/1.JBO.17.6.061214 JBOPFO 1083-3668 Google Scholar

43. 

J. Shah et al., “Photoacoustic imaging and temperature measurement for photothermal cancer therapy,” J. Biomed. Opt., 13 (3), 034024 (2008). http://dx.doi.org/10.1117/1.2940362 Google Scholar

44. 

M. Pramanik and L. V. Wang, “Thermoacoustic and photoacoustic sensing of temperature,” J. Biomed. Opt., 14 054024 (2009). http://dx.doi.org/10.1117/1.3247155 JBOPFO 1083-3668 Google Scholar

45. 

D. Hardie, A. Sangster and N. Cronin, “Coupled field analysis of heat flow in the near field of a microwave applicator for tumor ablation,” Electromagn. Biol. Med., 25 29 –43 (2006). http://dx.doi.org/10.1080/15368370600572953 Google Scholar

46. 

F. A. Duck, Physical Properties of Tissue: A Comprehensive Reference Book, Academic Press, London (1990). Google Scholar

47. 

H. Pennes, “Analysis of tissue and arterial blood temperatures in the resting human forearm,” J. Appl. Physiol., 1 (2), 93 –122 (1948). Google Scholar

48. 

W. L. Nyborg, “Solutions of the bio-heat transfer equation,” Phys. Med. Biol., 33 785 –792 (1988). http://dx.doi.org/10.1088/0031-9155/33/7/002 PHMBA7 0031-9155 Google Scholar

49. 

A. Mandelis and J. Vanniasinkam, “Theory of nonradiative decay dynamics in intensely pumped solid-state laser media via laser photothermal diagnostics,” J. Appl. Phys., 80 6107 –19 (1996). http://dx.doi.org/10.1063/1.363685 JAPIAU 0021-8979 Google Scholar

50. 

K. D. Cole et al., Heat Conduction using Green’s Functions, 2nd ed.CRC Press, Boca Raton (2011). Google Scholar

51. 

I. Kiricuta and V. Simplaceanu, “Tissue water-content and nuclear magnetic-resonance in normal and tumor tissues,” Cancer Res., 35 (5), 1164 –1167 (1975). Google Scholar

52. 

C. Gabriel, S. Gabriel and E. Corthout, “The dielectric properties of biological tissues: I. Literature survey,” Phys. Med. Biol., 41 2231 –2249 (1996). http://dx.doi.org/10.1088/0031-9155/41/11/001 PHMBA7 0031-9155 Google Scholar

53. 

S. Gabriel, R.W. Lau and C. Gabriel, “The dielectric properties of biological tissues. II. Measurements in the frequency range 10 Hz to 20 GHz,” Phys. Med. Biol., 41 2251 –2269 (1996). http://dx.doi.org/10.1088/0031-9155/41/11/002 PHMBA7 0031-9155 Google Scholar

54. 

S. Gabriel, R. W. Lau and C. Gabriel, “The dielectric properties of biological tissues. III. Parametric models for the dielectric spectrum of tissues,” Phys. Med. Biol., 41 2271 –2293 (1996). http://dx.doi.org/10.1088/0031-9155/41/11/003 PHMBA7 0031-9155 Google Scholar

55. 

W. T. Joines et al., “Microwave power absorption differences between normal and malignant tissue,” Int. J. Radiat. Oncol. Biol. Phys., 6 681 –687 (1980). http://dx.doi.org/10.1016/0360-3016(80)90223-0 Google Scholar

56. 

M. Lazebnik et al., “A large-scale study of the ultrawideband microwave dielectric properties of normal, benign and malignant breast tissues obtained from cancer surgeries,” Phys. Med. Biol., 52 6093 –6115 (2007). http://dx.doi.org/10.1088/0031-9155/52/20/002 PHMBA7 0031-9155 Google Scholar

57. 

R. Pethig, “Dielectric properties of body tissues,” Clin. Phy. Physiol. Meas., 8 (Suppl A), 5 –12 (1986). CPPMD5 0143-0815 Google Scholar

58. 

H. J. van Staveren et al., “Light scattering in Intralipid-10% in the wavelength range of 400–1100 nm,” Appl. Opt., 30 4507 –4514 (1991). http://dx.doi.org/10.1364/AO.30.004507 APOPAI 0003-6935 Google Scholar

Biography

Wei Wang received his bachelor’s degree at Xi'an Polytechnic University, China, his master’s degree from the University of Western Ontario, and his PhD from the University of Toronto. His research interests are PA imaging, optical imaging, ultrasound imaging, and thermal imaging.

Andreas Mandelis is a full professor of mechanical and industrial engineering and electrical and computer engineering at the Institute of Biomaterials and Biomedical Engineering, University of Toronto. He is the director of the Center for Advanced Diffusion-Wave and Photoacoustic Technologies (CADIPT), University of Toronto. His research interests include instrumentation and measurement science, nondestructive and noninvasive imaging and sensors, PA, photothermal and diffusion-wave physics and imaging technologies, signal processing, biothermophotonic laser diagnostics, nondestructive testing and imaging, and semiconductor optoelectronics.

© 2016 Society of Photo-Optical Instrumentation Engineers (SPIE) 1083-3668/2016/$25.00 © 2016 SPIE
Wei Wang and Andreas Mandelis "Microwave-heating-coupled photoacoustic radar for tissue diagnostic imaging," Journal of Biomedical Optics 21(6), 066018 (29 June 2016). https://doi.org/10.1117/1.JBO.21.6.066018
Published: 29 June 2016
Lens.org Logo
CITATIONS
Cited by 5 scholarly publications and 1 patent.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Tissues

Medium wave

Photoacoustic spectroscopy

Radar

Radar imaging

Tissue optics

Temperature metrology

Back to Top