Open access peer-reviewed chapter

Properties and Applications of Love Surface Waves in Seismology and Biosensors

Written By

Piotr Kiełczyński

Submitted: 08 May 2017 Reviewed: 15 February 2018 Published: 10 April 2018

DOI: 10.5772/intechopen.75479

From the Edited Volume

Surface Waves - New Trends and Developments

Edited by Farzad Ebrahimi

Chapter metrics overview

1,636 Chapter Downloads

View Full Metrics

Abstract

Shear horizontal (SH) surface waves of the Love type are elastic surface waves propagating in layered waveguides, in which surface layer is “slower” than the substrate. Love surface waves are of primary importance in geophysics and seismology, since most structural damages in the wake of earthquakes are attributed to the devastating SH motion inherent to the Love surface waves. On the other hand, Love surface waves found benign applications in biosensors used in biology, medicine, and chemistry. In this chapter, we briefly sketch a mathematical model for Love surface waves and present examples of the resulting dispersion curves for phase and group velocities, attenuation as well as the amplitude distribution as a function of the depth. We illustrate damages due to Love surface waves generated by earthquakes on real-life examples. In the following of this chapter, we present a number of representative examples for Love wave biosensors, which have been already used to DNA characterization, bacteria and virus detection, measurements of toxic substances, etc. We hope that the reader, after studying this chapter, will have a clear idea that deadly earthquakes and a beneficiary biosensor technology share the same physical phenomenon, which is the basis of a fascinating interdisciplinary research.

Keywords

  • Love waves
  • biosensors
  • earthquakes
  • surface acoustic waves
  • wireless sensors
  • dispersion curves

1. Introduction

It is interesting to note that many outstanding physicists (Kelvin, Michelson, and Jolly) expressed in the second half of the nineteenth century an opinion that classical physics (how we name it nowadays) is in principle completed and nothing interesting or significant rests to be discovered. Needless to say, forecasting development of future events was always and still is a very risky business, especially in physical sciences and engineering. Indeed, in these disciplines of human endeavors, one must take into account not only an inherently volatile human factor but also the impact of potential discoveries of unknown yet laws of nature, which often open new unanticipated possibilities and horizons. We may try to justify such an obvious complacency, attributed to the abovementioned scientists, by the historical spirit of the Belle Époque (1870–1914), that believed in harmony, good taste, optimism, unlimited progress and generally in positivistic philosophical ideas.

Anyway, not waiting for the revolution heralded by quantum mechanics (1900) or general theory of relativity (1917), classical physics was already shaken by the emergence of the theory of chaos (Poincaré 1882 and Hadamard 1898), which later on in the twentieth century will effectively eliminate deterministic description from many physical problems, such as weather forecasting, etc. Another new significant achievement of the classical physics (although not revolutionary) was the discovery of surface waves. At first, elastic surface waves were discovered in solids (Rayleigh 1885 and Love 1911) and then in electromagnetism (Zenneck 1907 and Sommerfeld 1909).

In fact, the existence of surface waves in solids was predicted mathematically by the celebrated British scientist Lord Rayleigh in 1885, who showed that elastic surface waves can propagate along a free surface of a semi-infinite body. By contrast to bulk waves, the amplitude of surface waves is confined to a narrow area adjacent to the guiding surface. Since surface waves are a type of guided waves, they can propagate often longer distances than their bulk counterparts and in addition, they are inherently sensitive to material properties in the vicinity of the guiding surface. It will be shown in the following of this chapter that these two properties of surface waves are of crucial importance in geophysics and sensor technology.

First, seismographs were constructed by British engineers in 1880, working in Japan for Meiji government. Consequently, the first long distance seismogram was registered in 1889 by German astronomer Ernst von Rebeur-Paschwitz in Potsdam (Germany), who was able to detect seismic signals generated by an earthquake occurred in Japan, some 9000 km away from Potsdam (Berlin). It was obvious soon that long distance seismograms display two different phases. First (preliminary tremor), a relatively weak signal arriving with the velocity of bulk waves (P and S) and second (main shock) with a much higher amplitude arriving with the velocity close to that of Rayleigh surface waves. However, this Rayleigh wave hypothesis was not satisfactory, since large part of the main shock energy was associated with the shear horizontal (SH) component of vibrations, absent by definition in Rayleigh surface waves composed of shear vertical (SV) and longitudinal (L) displacements. This dilemma was resolved in 1911 by the British physicist and mathematician Augustus Edward Hough Love by a brilliant stroke of thought [1]. Firstly, Love postulated that the SH component in the main shock is due to the arrival of a new type of surface waves (named later after his name) with only one SH component of vibrations. Secondly, Love assumed that SH surface waves are guided by an extra surface layer existing on the Earth’s surface, with properties different than those in the Earth’s interior. Using contemporary language, we can say that he made a direct hit.

It is noteworthy that the existence of Rayleigh and Love surface waves was first predicted mathematically prior to their experimental confirmation. This shows how beneficial can be the mutual interaction between the theory and experiment. Indeed, the theory indicates directions of future experimental research and the experiment confirms or renders the theory obsolete. It is worth noticing that the existence of a new type of electromagnetic surface waves was predicted mathematically quite recently, i.e., in 1988, and soon confirmed experimentally.

It is interesting to note that Love surface waves have direct counterparts in electromagnetism (optical planar waveguides) and quantum mechanics (particle motion in a quantum well). By contrast, a similar statement is not true for Rayleigh surface waves, which therefore remain a unique phenomenon within the frame of the classical theory of elasticity.

Surface waves of the Love type have a number of unique features. Firstly, they have only one SH component of vibrations. As a result, Love surface waves are insensitive to the loading with liquids of zero or negligible viscosities. Thus, Love surface waves can propagate long distances without a significant attenuation. Indeed, Love waves propagating many times around the Earth’s circumference have been observed experimentally. On the other hand, it was discovered much later (1981) that Love waves are very well suited for measurements of viscoelastic properties of liquids. Secondly, the mathematical description of Love surface waves is much simpler than that for Rayleigh surface waves. A relative simplicity of the mathematical model enables for direct physical insight in the process of Love wave propagation, attenuation, etc.

The idea to employ Love surface waves for measurements of viscoelastic properties of liquids was presented for the first time in 1981 by Kiełczyński and Płowiec in their Polish patent [2]. In 1987, the theory of the new method was presented by Kiełczyński and Pajewski on the international arena at the European Mechanics Colloquium 226 in Nottingham, UK [3]. In 1988, they presented this new method with equations and experimental results at the IEEE 1988 Ultrasonic Symposium in Chicago [4]. In 1989, Kiełczyński and Płowiec published a detailed theory and experimental results in the prestigious Journal of the Acoustical Society of America [5]. It is noteworthy that subsequent publications on Love wave sensors for liquid characterization appeared in USA not earlier than in 1992 [6], but nowadays, we witness about 100 publications per year on that subject [7].

We hope that the reader, after studying this chapter, will agree that the nature has many different faces and that the same physical phenomenon can be sometimes deadly (earthquakes) and in different circumstances, can be beneficiary (biosensor technology). As a consequence, SH surface waves of the Love type are an interesting example of an interdisciplinary research.

This chapter is organized as follows. Section 2 presents main characteristics and properties of Love surface waves, including basic mathematical model and examples of dispersion curves and amplitude distributions. More advanced mathematical treatment of the Love surface waves can be found, for example, in [8]. Section 3 shows the importance of Love surface waves in geophysics and seismology. Section 4 describes applications of Love surface waves in biosensors used in biology, medicine, chemistry, etc. Section 5 contains discussion of the chronological development of SH ultrasonic sensors starting from bulk wave sensors and then first surface wave sensors. We show also that the results of research conducted in Seismology and geophysics can be transferred to biosensor technology and vice versa. Conclusions and propositions for future research in biosensor technology employing Love surface waves are given in Section 6.

In addition to biosensors, Love surface waves are used in chemosensors, in non-destructive testing (NDT) of materials, and in sensors of various physical quantities such as:

  1. humidity of air [9];

  2. spatial distribution of elastic parameters in solid functionally graded materials (FGM) [10];

  3. elastic parameters of nanolayers [11];

  4. porosity of the medium [12]; and

  5. dielectric constant of liquids [13].

Recently, Love surface waves were also employed in the construction of the magnetic field sensor system with outstanding characteristics (sensitivity, dynamic range, etc.) [14].

Advertisement

2. Properties of Love surface waves

Shear horizontal (SH) surface waves of the Love type are elastic waves propagating in a surface waveguide, which is composed of a surface layer rigidly bonded to an elastic substrate, see Figure 1. The existence of an elastic surface layer is a necessary condition for propagation of Love surface waves, since it can be easily shown that on an elastic half-space alone, SH surface waves cannot exist. The extra surface layer must also be “slower” than the substrate, i.e., the following condition must hold [15]:

v1<v2E1

where v1 and v2 are phase velocities of bulk shear waves in the surface layer and substrate, respectively. In fact, the condition expressed by Eq. (1) allows for entrapment of partial waves in the surface layer due to the total reflection phenomenon occurring at the layer-substrate interface (x2=h). By contrast, if the condition given by Eq. (1) is not satisfied (v1>v2), then Love waves are evanescent in the direction of propagation x1 and, on average, no net power is transmitted along the surface waveguide.

Figure 1.

Basic structure of a free Love wave waveguide, not loaded with a viscoelastic liquid. An elastic surface layer of thickness “h” and a shear velocity v1 is rigidly bonded to the underlying semi-infinitive substrate with a shear velocity v2.

2.1. Dispersion equation of the Love surface wave

Mechanical displacement u3 of a time-harmonic Love surface wave propagating in the direction x1 has the following form:

u3x1x2t=fx2expjkx1ωtE2

where the function fx2 describes the amplitude of the Love wave as a function of the depth (x2 axis), k=ω/vp is the wavenumber of the Love wave, ω=2πf is its angular frequency, vp is the phase velocity of the Love wave and j=1. Since the surface waveguide is assumed to be lossless, the wavenumber k in Eq. (2) is a real quantity.

Substitution of Eq. (2) into Newton’s equation of motion leads to the Helmholtz differential equation for the transverse amplitude fx2. Solutions of the resulting Helmholtz differential equation have the following form [16]:

fx2=Acosq1x2cosq1h,for0x2<hsurface layerAeq2x2h,forhx2substrateE3

where

q12=k12k2q22=k22k2E4
k12=ω2v12k22=ω2v22E5

and A is an arbitrary constant. In isotropic solids, Love surface waves have two stress components, τ23andτ13, associated with the SH displacement u3. From Eq. (3), it follows that stress τ23 can be expressed by the following formula:

τ23x2=µ1,2fx2x2=µ1q1Asinq1x2cosq1h,for0x2<hsurface layerµ2q2Aeq2x2h,forhx2substrateE6

where μ1andμ2 are shear moduli of elasticity in the surface layer and substrate, respectively.

The mechanical displacement u3 and the associated stress τ23 must satisfy the appropriate boundary conditions, i.e., the continuity of u3 and τ23 at interfaces x2=0 (free guiding surface) and x2=h (the interface between the surface layer and the substrate). Substituting Eqs. (3) and (6) into the boundary conditions at x2=0 and x2=h, one obtains the following dispersion relation [16], for Love surface waves propagating in a planar waveguide shown in Figure 1:

Fωkω=μ1q1tanq1hμ2q2=0E7

Using Eq. (4), one can rewrite Eq. (7) in a more explicit form as:

Fωkω=μ11v121vp2tan2π1v121vp2fhμ21vp21v22=0E8

Eq. (8) shows that the unknown phase velocity vp of the Love surface wave is de facto an explicit function of the normalized product frequency-thickness fh, with v1andv2, being parameters. This property does not, however, hold for lossy Love wave waveguides where the elastic moduli μ1,μ2 as well as the velocities v1andv2 are implicit functions of the frequency f and are obviously independent of the surface layer thickness h.

The dispersion relation Eq. (8) is a transcendental algebraic equation for the unknown phase velocity vp and therefore can be solved only numerically using, for example, the Newton-Raphson iterative method [17].

2.2. Modal structure of the Love surface wave

The dispersion relation [Eq. (8)] reveals that phase velocity vp of the Love surface wave is a function of frequency. Hence, Love surface waves are dispersive. Moreover, since the function tangent in Eq. (8) is periodic, i.e., tanq1h=tanq1h+, where n=0,1,2,,etc., Love surface waves display a multimode structure.

The amplitude fx2 of the fundamental (n=0) mode of the Love surface wave, as a function of the distance x2from the guiding surface x2=0, is shown in Figure 2. It is clear that for sufficiently high frequencies, the energy of the Love wave is concentrated mostly in the surface layer in the vicinity of the guiding surface x2=0. By differentiation of Eq. (3), it is easy to show that the maximum of the amplitude fx2 occurs exactly at the free surface x2=0. By contrast, the associated stress τ23 vanishes at x2=0, i.e., at the free surface of the waveguide.

Figure 2.

Amplitude of the fundamental (n = 0) Love wave mode, as a function of the normalized depth x2/h, in a copper-steel waveguide, for different wave frequencies f = 3, 5, and 7 MHz, and surface layer thickness h = 100 μm.

2.3. Phase and group velocity of the Love surface wave

The total derivative of the implicit function Fωkω in the dispersion relation [Eq. (7)] with respect to the angular frequency ω equals:

ωFωkω+kFωkωdkω=0E9

Since group velocity vg of the Love surface wave, which describes the speed at which pulse envelope of the Love surface wave propagates, is defined as dk from Eq. 9, it is clear that:

vg=dk=kFωkωωFωkωE10

As a consequence, using Eqs. (7) and (10), one can show [8, 15, 16, 17, 18, 19] that group vg and phase vp velocities of the Love surface wave are connected via the following algebraic equation:

vgvpv22=μ1q2hsin2q1h2q1h+1+μ2cos2q1hμ1q2hv2v12sin2q1h2q1h+1+μ2cos2q1hE11

Eqs. (7) and (11) show that phase vp and group vg velocities of the Love surface wave in the low (f0) and high (f) frequency limits are the same and equal, respectively, v2 and v1.

The phase velocity resulting from the solution of Eq. (8) and the group velocity determined by Eq. (11) of the fundamental mode of Love surface waves, as a function of the normalized frequency fh, are given in Figure 3. From Figure 3, it is evident that for low frequencies, the phase and group velocities of Love surface waves approach asymptotically that of bulk shear waves v2 in the substrate. On the other hand, at high frequency limit, the phase and group velocities of the Love wave tend to the velocity v1, namely to the velocity of bulk shear waves in the surface layer.

Figure 3.

Phase vp and group vg velocities of the fundamental mode of the Love surface wave propagating in a copper-stainless steel waveguide, as a function of the normalized frequency-thickness product fh [MHz-mm], v1=2223.5m/s,andv2=3017m/s.

2.4. Influence of a viscoelastic liquid loading Love wave waveguides

It is noteworthy that in waveguides loaded with a lossy, viscoelastic liquid, the wavenumber k of the Love surface wave is a complex quantity, i.e., k=ω/vp+, where α is the coefficient of attenuation of the Love wave. Three most popular viscoelastic liquids are described by Kelvin-Voigt, Newton and Maxwell models, respectively [20]. The dispersion relation of Love surface waves propagating in waveguides loaded with a viscous liquid can be found in [21]. In lossy waveguides, the group velocity of Love waves cannot be rigorously defined [22]. As a result, the formula 11 is valid only approximately in lossy Love wave waveguides. As a matter of fact, in a waveguide loaded with a viscoelastic liquid, the amplitude of the Love wave is non-zero in a thin layer of the liquid adjacent to the surface layer of the waveguide. The penetration of the Love wave energy into the adjacent liquid is of crucial importance in understanding the operation of Love wave biosensors. Indeed, if Love wave energy was not penetrating in the measured liquid, the parameters of the Love wave might not be affected by the liquid and the operation of the whole sensor would be essentially impossible.

Advertisement

3. Love surface waves in seismology

Since Love surface waves were originally discovered in seismology, we give here a brief description of their applications in seismic and geophysical research.

Propagation of Love surface waves on the Earth’s surface is made possible by layered structure of the Earth. The outermost layer of the Earth, the crust, is made of solid rocks composed of lighter elements. Thickness of the crust varies from 5 to 10 km under oceans (oceanic crust) to 30–70 km under continents (continental crust). The crust sits on mantle, which in turn covers the outer and inner core. The destructive power of earthquakes is mainly due to waves traveling in this thin crustal layer [23].

As predicted by Love, the velocity of SH bulk waves increases with depth [24], i.e., as a function of distance from the free surface of the Earth.

The frequency of Love waves generated by earthquakes is rather low comparing to that used in sensor technology and ranges typically from 10 mHz to 10 Hz.

3.1. Investigation of the Earth’s interior with Love surface waves

Love and Rayleigh surface waves travel along great circle paths around the globe. Surface waves from strong earthquakes may travel several times around the Earth without a significant attenuation. They are termed global Rayleigh wave impulses [25]. An example of surface waves traveling multiply around the Earth [26] is given in Figure 4.

Figure 4.

Illustration of a seismogram of Rayleigh surface waves triggered by an earthquake. Note that, Rayleigh wave packet traveled 8 times around the Earth’s circumference.

Seismic waves, generated both by natural earthquakes and by man-made sources, have delivered an enormous amount of information about the Earth’s interior (subsurface properties of Earth’s crust). In classical seismology, Earth is modeled as a sequence of uniform horizontal layers (or spherical shells) having different elastic properties and one determines these properties from travel times and dispersion of seismic waves [27].

Love surface waves have been successfully employed in a tomographic reconstruction of the physical properties of Earth’s upper mantle [28] as well as in diamond, gold, and copper exploration in Australia, South America, and South Africa [29].

Surface waves generated by earthquakes or man-made explosions were used in quantitative recovery of Earth’s parameters as a function of depth. These seismic inverse problems helped to discover many fine details of the Earth’s interior [30, 31, 32].

It is noteworthy that many theoretical methods were initially originated in seismology and geophysics before their transfer to the surface wave sensor technology (see Table 1 in Section 5.5).

DevelopmentsSeismologyBiosensors
Basic theoryLove [1]Kiełczyński [3]
Multilayered waveguides (transfer matrix method)Haskell [72]Kiełczyński [8]
Viscoelastic waveguides (theoretical analysis)Sezawa [73]Kiełczyński [74]
Inverse problemsDorman [76]Kiełczyński [77]
Nonlinear wavesKalyanasundarm [78]
Phased arraysFrosch [79]
TomographyNakanishi [80]
Higher-order modesHaskell [81]
Solitary wavesBataille [82]
Energy harvestingQu [83]
Waveguides with nanomaterialsPenza [84]
Piezoelectric waveguidesKovacs [6]
ResonatorsKovacs [67]
Delay linesTournois [19]

Table 1.

Chronology of developments in Love wave biosensors and Love wave seismology.

3.2. Structural damages due to Love surface waves generated by earthquakes

An example of structural damages made by surface waves of the Love type is shown in Figure 5. It is apparent that railway tracks were deformed by strong shear horizontal SH forces parallel to the Earth’s surface. Love surface waves together with Rayleigh surface waves are the most devastating waves occurring during earthquakes.

Figure 5.

Twisted railroad tracks, an example of structural damages due to SH displacement of Love surface waves in the aftermath of an earthquake.

3.3. Application of metamaterials to minimize devastating effects of Love surface waves in the aftermath of earthquakes

It is interesting to note that recently developed earthquake engineered metamaterials open a new way to counterattack seismic waves [33, 34]. The metamaterials actively control the seismic waves by providing an additional shield around the protected building rather than reconstructing the building structure. Compared with common engineering solutions, the advantage of the metamaterial method is that it can not only attenuate seismic waves before they reach critical targets, but also protect a distributed area rather than an individual building. The periodic arrangement of metamaterial structure creates frequency band gaps, which effectively prevent surface waves propagation on the Earth’s surface via a Bragg scattering mechanism.

Advertisement

4. Biosensors employing Love surface waves

A biosensor can be described as a device which can generate a signal (usually electrical) that is proportional to the concentration of a particular biomaterial or chemicals in the presence of a number of interfering species [35]. This can be accomplished using biological recognition elements such as enzymes, antibodies, receptors, tissues, and microorganisms as sensitive materials because of their selective functionality for target analytes along with an appropriate transducer.

4.1. Confinement of the energy of Love surface waves near the free surface of the waveguide

High sensitivity of Love surface wave sensors can be explained by spatial concentration of the energy of Love waves. Indeed, it was shown in Section 2 that the energy of Love surface waves is localized mostly in the vicinity of the free guiding surface (Figure 1), looking in both sides from it. Moreover, the amplitude of Love surface waves reaches maximum at the free guiding surface x2=0 (Figure 2). Therefore, we can expect that propagation of Love surface waves will be to a lesser or higher extent perturbed by a material (such as liquid) being in contact with the guiding surface. This feature of Love waves was exploited in the construction of various biosensors used for detection and quantification of many important parameters of biological and chemical substances [21, 36, 37, 38, 39, 40].

4.2. Correlation between concentration of the measured analyte and parameters of the Love surface wave

Surface waves of the Love type are especially suited to measure parameters of viscoelastic liquids, polymers, gels, etc., providing that they can form a good mechanical contact (absorption and adhesion) with free surface of the waveguide. Since Love surface waves are, in principle, mechanical waves, they can measure the following mechanical parameters of an adjacent medium: density, modulus of elasticity, and viscosity. In waveguides composed of piezoelectric elements (substrate and/or surface layer), dielectric constant of the adjacent medium will also affect the propagation of Love surface waves. In practice, we are interested in detection and quantification other more specific properties of biological and chemical materials, such as concentration and presence of proteins, antibodies, toxins, bacteria, viruses, size and shape of DNA, etc. Therefore, the next step in the development of Love wave sensors is to correlate (experimentally or analytically) the abovementioned specific properties of the measured analytes with changes in density, viscosity, and elastic moduli of the surface (sensing) layer. Finally, we have to measure changes in phase velocity and attenuation of Love surface waves, which are due to changes in density, viscosity, and elastic moduli of this surface layer. It should be noticed that part of Love wave energy enters into the measured liquid to some distance (penetration depth) from the guiding surface. Such an energy redistribution changes certainly the phase velocity and attenuation of the Love surface wave. In practice, we often adopt a more empirical approach, i.e., we measure directly changes in phase velocity and attenuation of Love waves, as a function of the aforementioned specific properties of the measured material, such as the concentration of proteins and so on, without referring to changes in density, viscosity or elastic modulus of the measured material. However, the former step is indispensable during modeling, design, and optimization of Love surface wave sensors.

4.3. Parameters of the Love surface wave measured

As with other types of wave motion, we can measure in principle two parameters of Love surface waves, i.e., their phase and amplitude. Polarization of SH surface waves of the Love type is constant and therefore does not provide any additional information about the medium of propagation. Phase Φx1t measurements in radians are directly related to the phase velocity vp of Love surface waves via the following equation:

Φx1t=kx1ωt=ωx1vptE12

Similarly, amplitude A measurements are correlated with the coefficient of attenuation α (in Np/m) of Love surface waves as follows:

α=1x2x1lnA1A2Np/mE13

where A1 and A2 are two amplitudes of the wave measured at points x1andx2, respectively (x2>x1). In order to obtain the coefficient of attenuation in dB/m, the coefficient α given by Eq. (13) must be multiplied by 20loge8.686.

4.4. Sensors working in a resonator and delay line configurations

Phase and amplitude characteristics of Love surface waves can be measured in a closed loop configuration by placing Love wave delay line in a feedback circuit of an electrical oscillator (resonator). Another possibility is to use network analyzer, which provides phase shift and insertion loss of the Love wave sensor working in an open loop configuration, due to the load of the sensor with a measured material. The typical frequency range used by Love wave sensors is from 50 MHz to 500 MHz [7].

The structure and cross section of a typical Love wave biosensor is shown in Figure 6a and b. A relatively thick (0.5–1.0 mm) substrate provides mechanical support for the whole sensor. Often the substrate material is piezoelectric (AT-cut quartz material [41]). In this case, a pair of interdigital transducers (IDTs) can be deposited on the substrate to form a delay line of the sensor. The guiding layer (SiO2, ZnO, PMMA, etc.), deposited directly on the substrate, provides entrapment for surface wave energy. The sensing layer, made of gold (Au) or a polymer, usually very thin (˜50-100 nm), serves as an immobilization area for the measured biological material. This thin-sensing layer interacts directly with the measured material (liquid) and serves often as a selector of the specific target substance, such as antigen, to be measured.

Figure 6.

a) Layered structure of a typical Love wave sensor not yet connected to the external driving circuit and b) cross-section of this sensor structure + loading liquid.

4.5. Sensors controlled remotely by wireless devices

An interesting solution for Love wave sensors was proposed in [42], where the Love wave sensor works in a wireless configuration without an external power supply. This design has many unique advantages, i.e., the sensor can be permanently implanted in a patient body to monitor continuously the selected property of a biological liquid. Readings of the sensor can be made on demand, totally noninvasively by a reading device connected to a broader computer system of patient monitoring. Another implementation of a remotely controlled wireless Love wave sensor was presented in [43]. The proposed sensor can measure simultaneously two different analytes using Love surface waves with a frequency of 440 MHz.

Wireless bioelectronics sensors may be used in a variety of fields including: healthcare, environmental monitoring, food quality control, and defense.

4.6. Examples of laboratory and industrial grade Love wave sensors

To apply the measured analyte to the Love wave sensor, the sensor is often equipped with a flow cell, which separates interdigital transducers from sensing area of the waveguide [44]. A laboratory grade Love wave sensor equipped with a flow cell is shown in Figure 7.

Figure 7.

An example of a laboratory grade Love wave sensor with a flow cell [42].

A prototype of an commercial ready Love wave sensor was presented in 2015 in Ref. [45]. A 250 MHz delay line Love wave immunosensor was designed on the ST quartz substrate with a thin gold layer of thickness ˜90 nm used as a guiding and sensing area, for antibodies or antigens can be easily immobilized on a gold surface. The changes of Love wave velocity and attenuation were due to antibodies-antigens interactions. A disposable test cassette with embedded Love wave immunosensor is connected to a handheld electronic reader, which in turn is connected wirelessly via bluetooth to a smartphone or a computer. This device is a strong candidate for clinical and personnel healthcare applications.

4.7. Examples of analytes measured by Love wave biosensors

Love wave biosensors have been used in measurement and detection of a large number of substances (analytes) [44]. As representative examples, we can mention the following:

  • concentration of bovine serum albumin [46];

  • real-time detection of antigen-antibody interactions in liquids (immunosensor) [47];

  • simultaneous detection of Legionella and E. coli bacteria [48];

  • virus and bacteria detection in liquids [49];

  • detection of pathogenic spores Bacillus anthracis below inhalation infectious levels [50];

  • investigation of lipid specificity of human antimicrobial peptides [51];

  • Sin Nombre Virus detection at levels lower than those typical for human patients suffering from hantavirus cardiopulmonary syndrome [52];

  • detection of nanoparticles in liquid media [53];

  • okadaic acid detection [54];

  • study of protein layers [55];

  • antibody binding detection [56];

  • toxicity of heavy metals [57];

  • size and shape of DNA [58];

  • real-time detection of hepatitis B [59];

  • liquid chromatography [60];

  • immunosensors for detection of pesticide residues and metabolites in fruit juices [61];

  • detection of cocaine [62]; and

  • detection of carbaryl pesticide [63].

4.8. Desired characteristics (features) of industrial grade Love wave sensors

This rather impressive list of achievements in R&D activities on biosensor technology suggests that biosensors employing Love surface waves have a huge potential. However, in order to compete with other types of biosensors, such as optical sensors based on the surface plasmon resonance [64], the biosensors employing Love surface waves should possess the following characteristics:

  • high sensitivity to the measured property (measurand);

  • high selectivity to the measured property (measurand);

  • low limit of detection;

  • zero temperature coefficient (high-thermal stability);

  • high repeatability and stability;

  • possibility of multiple reuse; and

  • cost-effectiveness.

At present, none of the above targets have been fully achieved. Love wave biosensors are, in general, still in the laboratory research phase, where most developments are focused on the proof of concept and construction of a working prototype. Only one European company offers today commercially available Love wave sensors [7]. Nevertheless, as it was shown in this section, Love wave biosensors can be used to measurements of a surprisingly large number of biological substances (analytes) with a quite remarkable accuracy and sensitivity. Therefore, in our opinion, Love wave biosensors will reach soon an industrial grade level with numerous real-life applications in biology, medicine (clinical practice), and chemistry.

Advertisement

5. Discussion

5.1. Older sensors using bulk SH waves

It is interesting to note that first acoustic sensors for measurements of viscoelastic properties of liquids used to this end bulk (not surface) SH waves propagating in a solid buffer, loaded on one side with a measured viscoelastic liquid. This idea appeared in 1950 in works of such prominent ultrasonic scientists Mason and McSkimmin [65]. However, the main drawback of the bulk wave sensors was their inherent low sensitivity. For example, to perform measurements with a water-loaded sensor, one had to observe about 50 consecutive reflections in the solid buffer.

5.2. Emergence of new sensors using SH surface waves of the Love type

The breakthrough came with a proposition to employ to this end SH surface waves of the Love and Bleustein-Gulyaev types. This idea was first articulated by Kiełczyński and Płowiec in 1981 in their Polish patent no 130040 [2]. In 1987, the theory of the new method was presented by Kiełczyński and Pajewski on the international arena at the European Mechanics Colloquium 226 in Nottingham, UK [3]. In 1988, this new method, with equations and experimental results, was presented by Kiełczyński and Pajewski at IEEE 1988 Ultrasonic Symposium in Chicago [4]. In 1989, Kiełczyński and Płowiec published detailed theory and experimental results in the prestigious Journal of the Acoustical Society of America [5]. Their theory [3, 4, 5] was based on the Auld’s perturbative technique [66] and gave satisfactory results for liquids of viscosities up to ˜10 Pas. The main advantage of the Love surface wave sensors is their very high sensitivity, namely the sensitivity of a few orders of magnitude (102 to 104) higher than that of their bulk SH waves counterparts [3, 4, 5]. As a result, measurements of the viscosity of water (˜1 mPas) and other biological substances (based largely on water) was no longer a challenge, what was the case with bulk SH wave sensors. In other words, due to the employment of SH surface waves, the way for development of the corresponding biosensors was widely open.

It should be noticed that next publications on the Love wave sensors for liquid characterization appeared in the open literature not earlier than in 1992 [6]. In fact, in papers published in 1992, Kovacs and Venema [67], and, in 1993, Gizeli et al. [68] confirmed our earlier discovery [3, 4, 5] that Love surface waves are much more sensitive to viscous loading than other types of SH waves. In another paper published in 1992, Gizeli et al. [69] developed theoretical analysis for Love wave sensors, using the same Auld’s perturbative technique [66] as that employed by us in papers [3, 4, 5].

It is interesting to note that two other types of SH waves, i.e., leaky SH SAW waves and plate SH waves, were also tried to measure viscosity of liquids. Leaky SH SAW waves were proposed in 1987 [70] by Moriizumi et al. and SH plate waves in 1988 by Martin et al. [71]. However, these two types of SH waves were quickly abandoned, since the corresponding viscosity sensors were of inherently low sensitivity, difficult in practical realization and difficult in theoretical analysis (leaky SH SAW waves). In fact, the energy of SH plate waves is uniformly distributed across the whole thickness of the plate. Therefore, SH plate waves are not so sensitive to viscous loading as Love surface waves, whose energy is highly concentrated in the surface layer of the waveguide. On the other hand, leaky SH SAW waves are not pure SH waves and contains in principle all three components of vibrations, not only the SH one. In particular, the component perpendicular to free surface of the waveguide will continuously radiate energy into the adjacent liquid. This will cause an additional attenuation for leaky SH SAW waves, which will be indistinguishable from that due to the viscous loading measured.

5.3. Mathematical apparatus and numerical methods used in analysis of Love surface waves

R&D activities in seismology and biosensor technology using Love surface waves focus inevitably on different problems and challenges. The main reason for these differences is the nature and scale of Love surface waves used in seismology and biosensor technology, i.e., in seismology, they are a natural phenomenon and in biosensors, they are controlled within man-made devices. It is instructive to compare the chronology of developments made in seismology and in biosensor technology (see Table 1). In fact, the theory of Love waves published in 1911 [1] was developed for the simplest surface wave waveguide, namely for that composed of linear, isotropic, and lossless materials (surface layer on a substrate). Since loading viscoelastic liquids are always lossy, the corresponding theory of Love wave sensors had to use perturbative [3] or numerical methods [37]. The theory of Love waves in multilayered waveguides, developed in Seismology [72], uses a conventional transfer-matrix method based on the elementary matrix algebra. By contrast, the theory developed for biosensors extends the transfer-matrix method to a more advanced formalism of matrix differential equations with eigenvectors and eigenvalues and operator functions [8]. First theories of Love waves propagating in viscoelastic waveguides, were developed in Seismology [73], long before the advent of modern fast digital computers. By contrast, the corresponding theory developed for biosensors [74] in 2016 heavily relates on numerical methods.

5.4. Milestones in developments of Love wave seismology and Love wave biosensors

Examination of Table 1 reveals that a number of R&D activities already well established in Seismology were not yet initiated in biosensor technology. As examples, one can mention the applications of nonlinear Love waves, higher-order Love wave modes or solitary waves. This suggests that in future research, it may be advantageous to employ higher-order modes, nonlinear Love waves, metamaterials, etc., to increase biosensors sensitivity [75] or lower their limit of detection. Other technologies not yet used in biosensor technology are phased array and tomography. Indeed, applied to biosensors they may allow for a 2D characterization of the analyte distribution, electronic beam steering, focusing, etc. These indications for future research in biosensor technology show clearly advantages of multidisciplinary R&D activities, in this case seismology and biosensor technology. Indeed, it is much easier to adapt an existing technology already developed in other fields to a new domain than to invent a new technology from scratch without any prior feedback.

5.5. Novelty of the present chapter

Despite the fact that the first theory of Love surface waves was published as early as in 1911 [1], surprisingly, a large number of problems concerning the theory of Love surface waves have not yet been solved.

This chapter contains theoretical foundations and calculation results regarding the propagation of the Love wave in various media. A new interpretation of the Love wave dispersion equation was given. This equation is presented in the form of an implicit function of two variables, i.e., (ω,k). This allowed to evaluate the analytical dependencies on group velocity of Love wave propagating in a wide class of layered waveguides, e.g., in graded waveguides. This problem will be the subject of future author’s works.

The obtained results can be employed in the design and optimization of not only biosensors but also chemosensors and sensors of physical quantities that use Love waves. In addition, the obtained results can be used in seismology and geophysics for the interpretation of seismograms and determining the distribution of elastic parameters of the Earth’s crust.

This chapter contains also a novel comparison of milestones in developments made in Love wave seismology and Love wave biosensors (see Section 5.4). Since Love wave biosensors appeared exactly 70 years [2] after emergence of Love surface waves in seismology [1], it is not surprising that many discoveries and developments were made first in seismology and then transferred to biosensors (see Table 1). This cross-pollination between the two seemingly distant branches of science is very beneficial and can significantly accelerate developments made in either of them.

Advertisement

6. Conclusions

In this limited space chapter, it was impossible to address or even mention all interesting problems relevant to the properties and applications of Love surface waves in seismology and biosensor technology. Instead, we tried to present only main properties of the Love surface waves, such as their dispersive nature, phase and group velocities, amplitude distribution, etc., as well as their most iconic applications in seismology and biosensor technology. We think that presentation of the Love surface waves R&D activities in a broader historical perspective gives an invaluable insight in the process of developments made in this fascinating interdisciplinary domain of research.

In this chapter, we attempted to present a variety of aspects that can be attributed to SH surface waves of the Love type. As a matter of fact, Dr. Jekyll and Mr. Hyde Love surface waves possess simultaneously two diametrically different faces, i.e., first benign (biosensors) and second deadly (earthquakes). The good news is that developments made in one of these domains can be easily transferred to the second one and vice versa. In fact, Love surface waves were first discovered in seismology (1911). They finally enabled for precise interpretation of seismograms registered in the aftermath of earthquakes. Beneficiary applications (biosensors) of Love surface waves were announced exactly 70 years later (1981) in a Polish patent.

Since earthquake is a natural phenomenon, we have little or no influence on its occurrence and dynamics. By contrast, the construction and the operation of biosensors can be optimized by mathematical modeling and experimental studies. At present, the mathematical modeling of Love wave biosensors is an active domain of research. On the other hand, progress in electronics and computer technology will lead to development of new compact and reliable instrumentation working in conjunction with Love wave biosensors.

Despite their centennial heritage, Love surface waves are subject of an intensive research activity. For example, one can mention the application of inverse problem techniques to recover material parameters of surface layers from measurements of velocity and attenuation of Love surface waves. Inverse problem techniques have been successfully employed in seismology and geophysics [25] and recently also pioneered by the authors [74, 77, 85] and others [86, 87] in the biosensor technology.

Other open problems in the theory and technique of Love surface waves are non-linear Love waves, extremely slow Love waves [88], Love waves in layered nanostructures [89], energy harvesting with Love waves, and metamaterial-based seismic shielding, [33, 34], etc.

Finally, coming back to the idea expressed at the beginning of the introduction in this chapter, we want to assure the reader that there exist still many significant unresolved problems in the theory and technique of the Love surface waves, which deserve to be addressed in future R&D activities. We hope that this chapter may be helpful in this endeavor.

References

  1. 1. Love AEH. Some Problems of Geodynamics. UK: Cambridge University Press; 1911. pp. 89-104 and 149-152
  2. 2. Kiełczyński P, Płowiec R. Polish patent no. 130040; 1981
  3. 3. Kiełczyński P, Pajewski W. Determination of the rheological parameters of viscoelastic liquids using shear surface waves. In: Parker DF, Maugin GA, editors. Recent Developments in Surface Acoustic Waves. In: Proceedings of European Mechanics Colloquium 226; University of Nottingham, UK; Springer Series on Wave Phenomena 7; September 2-5, 1987. Berlin: Springer-Verlag; 1988. pp. 317-321
  4. 4. Kiełczyński P, Pajewski W. A new method for the determination of the shear impedance of viscoelastic liquids. IEEE1988 Ultrasonic Symposium Proceedings; USA. 1988. pp. 323-326
  5. 5. Kiełczyński P, Płowiec R. Determination of the shear impedance of viscoelastic liquids using Love and Bleustein-Gulyaev surface waves. Journal of Acoustical Society of America. August 1989;86(2):818-827
  6. 6. Kovacs G, Lubking GW, Vellekoop MJ, Venema A. Love waves for (bio)chemical sensing in liquids. In: Ultrasonics Symposium Proceedings; IEEE; Tucson, USA. 1992. pp. 281-285
  7. 7. Gronewold TMA. Surface acoustic wave sensors in the bioanalytical field: Recent trends and challenges. Analytica Chimica Acta. 2007;603:119-128
  8. 8. Kiełczyński P, Szalewski M, Balcerzak A, Wieja K. Propagation of ultrasonic love wave in non-homogeneous elastic functionally graded materials. Ultrasonics. 2016;65:220-227
  9. 9. Lan XD, Zhang SY, Fan L, Wang Y. Simulation of SAW humidity sensors based on (112¯0) ZnO/R-sapphire structures. Sensors. 2016;16:1112. DOI: 10.3390/s16111112
  10. 10. Kiełczyński P, Szalewski M, Balcerzak A, Wieja K. Group and phase velocity of Love waves propagating in elastic functionally graded materials. Archives of Acoustics. 2015;40:273-281
  11. 11. Zhang S, Gu B, Zhang H, Feng XQ, Pan R, Alamusi, Hu N. Propagation of Love waves with surface effects in an electrically-shorted piezoelectric nanofilm on a half-space elastic substrate. Ultrasonics. 2016;66:65-71
  12. 12. Wang YS, Zhang ZM. Propagation of Love waves in a transversely isotropic fluid-saturated porous layered half-space. Journal of the Acoustical Society of America. February 1998;103(2):695-701
  13. 13. Xia Q, Chen Z, Wang M. The system design of a love wave sensor for measuring liquid dielectric constant. IEEE International Ultrasonics Symposium (IUS); 18-21 October 2011; Orlando, FL, USA. DOI: 10.1109/ULTSYM.2011.0571
  14. 14. Kittmann A, Durdaut P, Zabel S, Reermann J, Schmalz J, Spetzler B, Meyners D, Sun NX, McCord J, Gerken M, Schmidt G, Höft M, Knöchel R, Faupel F, Quandt E. Wide band low noise love wave magnetic field sensor system. Scientific Reports. 2018;8:278. DOI: 10.1038/s41598-017-18441-4
  15. 15. Auld BA. Acoustics Fields and Waves in Solids, Vol. II. Florida: Krieger Publishing Company; 1990. p. 95
  16. 16. Royer D, Dieulesaint E. Elastic Waves in Solids I. Berlin: Springer; 2000. p. 306
  17. 17. Press WH, Teukolsky SA, Vetterling WT, Flannery BP. Numerical Recipies. 3rd ed. New York: Cambridge Univeristy Press; 2007. p. 456
  18. 18. Kiełczyński P, Szalewski M, Balcerzak A, Wieja K. Group and phase velocities of Love waves propagating in elastic functionally graded materials. Archives of Acoustics. 2015;40(2):273-281
  19. 19. Tournois P, Lardat C. Love wave-dispersive delay lines for wide-band pulse compression. IEEE Transactions on Sonics and Ultrasonics. 1969;16(3):107-117
  20. 20. Christensen RM. Theory of Viscoelasticity. 2nd ed. Academic Press; 1982
  21. 21. Kiełczyński P, Szalewski M, Balcerzak A. Effect of a viscous loading on Love wave propagation. International Journal of Solids and Structures. 2012;49:2314-2319
  22. 22. Auld BA. Acoustics Fields and Waves in Solids, Vol. I. Florida: Krieger Publishing Company; 1990. p. 227
  23. 23. Fowler CMR. The Solid Earth: An Introduction to Global Geophysics. 2nd ed. The Edinburgh Building, Cambridge, UK: Cambridge University Press; 2005
  24. 24. http://www.tulane.edu/˜sanelson/eens1110/earthint.htm
  25. 25. Bormann P, Engdahl ER, Kind R. Chapter 2: Seismic wave propagation and earth models. IASPEI New Manual of Seismological Observatory Practice. ftp://ftp.ingv.it/pub/elisabetta.danastasio/grot/manual_seismological_observatory-2002.pdf
  26. 26. Romanowicz B. Inversion of Surface Waves, A Review, International Handbook of Earthquake & Engineering Seismology, Part A, Vol. 81A. 1st ed. Part II. 2002. p. 149
  27. 27. Haney M, Douma H. Inversion of Love wave phase velocity, group velocity and shear stress ratio using finite elements. SEG San Antonio 2011 Annual Meeting
  28. 28. Foster A, Ekstrom G, Nettles M. Surface wave phase velocities of western unites states from a two-station method. Geophysical Journal International. 2014;196(2):1189-1206
  29. 29. Fishwick S, Rowlinson N. 3-D structure of the Australian lithosphere from evolving seismic datasets. Australian Journal of Earth Sciences. 2012;59:809-826
  30. 30. The seismic reflection inverse problem. Inverse Problems, Vol. 25(12). IOP Publishing Ltd; 2009. WW Symes Published
  31. 31. Haney MM, Tsai VC. Nonperturbational surface-wave inversion: A Dix-type relation for surface waves. Geophysics. November-December 2015;80(6):EN167-EN177
  32. 32. Lai CK, Wilmanski K, editors. Surface Waves in Geomechanics: Direct and Inverse Modelling for Soils and Rocks. New York: Springer; 2005
  33. 33. Qiujiao D, Yi Z, Huang G, Yang H. Elastic metamaterial-based seismic shield for both lamb and surface waves. AIP Advances. 2017;7:075015
  34. 34. Palermo A et al. Engineered metabarrier as shield from seismic surface waves. Scientific Reports. 2016;6:39356. DOI: 10.1038/srep39356
  35. 35. Thévenot DR, Toth K, Durst RA, Wilsond GS. Electrochemical biosensor: Recommended definitions and classification. Biosensors and Bioelectronics. 2001;16:121-131
  36. 36. Tada K, Nozawa T, Kondoh J. Real-time monitoring of methanol concentration using a shear horizontal surface acoustic wave sensor for direct methanol fuel cell without reference liquid measurement. Japanese Journal of Applied Physics. 2017;56:07JD15
  37. 37. Kiełczyński P. Attenuation of Love waves in low-loss media. Journal of Applied Physics. 1997;82:5932-5937
  38. 38. Caliendo C, Sait S, Boubenider F. Love-mode MEMS devices for sensing applications in liquids. Micromachines. 2016;7:15
  39. 39. Wang W, He S. Sensitivity evaluation of a Love wave sensor with multi-guiding-layer structure for biochemical application. Sensors & Transducers Journal. September 2008;96(9):32-41
  40. 40. Li S, Wan Y, Fan C, Su Y. Theoretical study of monolayer and double-layer waveguide Love wave sensors for achieving high sensitivity. Sensors. 2017;17:653
  41. 41. Royer D, Dieulesaint E. Elastic Waves in Solids I. Berlin: Springer; 2000. p. 207
  42. 42. Wang W, He S. A Love wave reflective delay line with polymer guiding layer for wireless sensor application. Sensors. 2008;8:7917-7929
  43. 43. Haekwan O, Chen F, Kim K, Lee K. Wireless and simultaneous detections of multiple bio-molecules in a single sensor using Love wave biosensor. Sensors. 2014;14:21660-21675
  44. 44. Gaso MIR, Jiménez Y, Francis LA, Arnau A. Love wave biosensors: A review, state of the art in biosensors—General aspects, Chapter 11. In: Rinken T, editor. InTech, Chapters published March 13, 2013 under CC BY 3.0 license. DOI: 10.5772/45832 Edited Volume, Open Access; ISBN 978-953-51-1004-0, 360 pages
  45. 45. Yatsuda H, Kano K, Kogai T, Yoshimura N, Goto M. Rapid diagnostic test using SH-SAW immunosensor. In: Six International Symposium on Acoustic Wave Devices for Future Mobile Communication Systems; November 24-25 2015; Chiba University, Japan
  46. 46. Goto M, Yatsuda H, Kondoh J. Effect of viscoelastic film for shear horizontal surface acoustic wave on quartz. Japanese Journal of Applied Physics. 2015;54:07HD02
  47. 47. Harding GL, Du J, Dencher PR, Barnett D, Howe E. Love wave acoustic immunosensor operation in liquid. Sensors and Actuators, A: Physical. 1997;61(1-3):279-286
  48. 48. Howe E, Harding G. A comparison of protocols for the optimisation of detection of bacteria using a surface acoustic wave (SAW) biosensor. Biosensors & Bioelectronics. 2000;15(11-12):641-649
  49. 49. Tamarin O, Comeau S, Déjous C, Moynet D, Rebière D, Bezian J, Pistré J. Real time device for biosensing: Design of a bacteriophage model using Love acoustic wave. Biosensors & Bioelectronics. 2003;18:755-763
  50. 50. Branch DW, Brozik SM. Low-level detection of a bacillus anthracis simulant using Love-wave biosensors on 36° YX LiTaO3. Biosensors & Bioelectronics. 2004;19:849-848
  51. 51. Andrä J, Böhling A, Gronewold TMA, Schlecht U, Perpeet M, Gutsmann T. Surface acoustic wave biosensor as a tool to study the interactions of antimicrobial peptides with phospholipid and lipopolysaccharide model membranes. Langmuir. 2008;24(16):9148-9153
  52. 52. Bisoffi M, Hjelle B, Brown DC, Branch DW, Edwards TL, Brozik SM, Bondu-Hawkins VS, Larson RS. Detection of viral bioagents using a shear horizontal surface acoustic wave biosensor. Biosensors and Bioelectronics. 2008;23(9):1397-1403
  53. 53. El Fissi L, Friedt J-M, Luzet V, Chérioux F, Martin G, Ballandras S. A Love-wave sensor for direct detection of biofunctionalized nanoparticles. In: Proceedings from the IEEE frequency Control Symposium. 2009:861-865
  54. 54. Fournel F, Baco E, Mamani-Matsuda M, Degueil M, Bennetau B, Moynet D, Mossalayi D, Vellutini L, Pillot J-P, Dejous C, Rebiere D. Love wave biosensor for real-time detection of okadaic acid as DSP phycotoxin. In: Proceedings of the Eurosensors XXIV. Austria: Linz. p. 2010
  55. 55. Saha K, Bender F, Rasmusson A, Gizeli E. Probing the viscoelasticity and mass of a surface-bound protein layer with an acoustic waveguide device. Langmuir. 2003;19:1304-1311
  56. 56. Kardous F, El Fissi L, Friedt J-M, Bastien F, Boireau W, Yahiaoui R, Manceau JF, Ballandras S. Integrated active mixing and biosensing using low frequency vibration mixer and Love-wave sensor for real time detection of antibody binding event. Journal of Applied Physics. 2011;109:1-8
  57. 57. Gammoudi I, Tarbague H, Lachaud JL, Destor S, Othmane A, Moyenet D, Kalfat R, Rebiere D, Dejous C. Love wave bacterial biosensors and microfluidic network for detection of heavy metal toxicity. Sensor Letters. 2011;9(2):816-819
  58. 58. Tsortos A, Papadakis G, Mitsakakis K, Melzak KA, Gizeli E. Quantitative determination of size and shape of surface-bound DNA using an acoustic wave sensor. Biophysical Journal. 2008;94:2706-2715
  59. 59. Lee HJ, Namkoong K, Cho EC, Ko C, Park JC, Lee SS. Surface acoustic wave immunosensor for real-time detection of hepatitis B surface antibodies in whole blood samples. Biosensors and Bioelectronics. June 2009;24(10):3120-3125
  60. 60. Gizeli E, Stevenson AC, Goddard NJ, Lowe CR. A novel Love-plate acoustic sensor utilizing polymer overlayers. IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control. 1992;39(5):657-659
  61. 61. March C, Manclús JJ, Jiménez Y, Arnau A, Montoya A. A piezoelectric immunosensor for the determination of pesticide residues and metabolites in fruit juices. Talanta. 2009;78(3):827-833
  62. 62. Stubbs DD, Lee S-H, Hunt WD. Investigation of cocaine plumes using surface acoustic wave immunoassay sensors. Analytical Chemistry. 2003;75(22):6231-6235
  63. 63. Rocha-Gaso M-I, García J-V, García P, March-Iborra C, Jiménez Y, Francis L-A, Montoya Á, Arnau A. Love wave Immunosensor for the detection of carbaryl pesticide. Sensors. 2014 September;14(9):16434-16453
  64. 64. In: Rinken T, editor. State of the Art in Biosensors—General Aspects. 2013. Open Access. DOI: 10.5772/45832
  65. 65. Moore RS, McSkimin HJ. Dynamic shear properties of solvents and polystyrene solutions from 20 to 300 MHz. In: Mason WP, Thurston RN, editors. Physical Acoustics VI. New York: Academic Press; 1970. p. 167
  66. 66. Auld BA. Acoustics Fields and Waves in Solids, Vol. II. Florida: Krieger Publishing Company; 1990. p. 275
  67. 67. Kovacs G, Venema A. Theoretical comparison of sensitivities of acoustic shear wave modes for (bio)chemical sensing in liquids. Applied Physics Letters. 1992;61:639-641
  68. 68. Gizeli E, Stevenson AC, Goddard NJ, Lowe CR. Acoustic Love plate sensors: Comparison with other acoustic devices utilizing surface SH waves. Sensors and Actuators B. 1993;13-14:638639
  69. 69. Gizeli E, Goddard NJ, Lowe CR, Stevenson AC. A Love plate biosensor utilising a polymer layer. Sensors and Actuators B. 1992;6:131-137
  70. 70. Moriizumi T, Unno Y, Shiokawa S. New sensor in liquid using leaky SAW. Proceedings of the Ultrasonics Symposium of IEEE. 1987:579-582
  71. 71. Martin SJ, Ricco AJ, Frye GC. Sensing in Liquids with SH Plate Mode Devices. 1988 Ultrasonics Symposium Proceedings; 1988. pp. 607-611
  72. 72. Haskell NA. The dispersion of surface waves in multilayered media. Bulletin of the Seismological Society of America. 1953;43:17-34
  73. 73. Sezawa K. On the decay of waves in visco-elastic solid bodies. Bulletin of the Earthquake Research Institute (Japan). 1927;3:43-53
  74. 74. Kiełczyński P, Szalewski M, Balcerzak A, Wieja K. Evaluation of viscoelastic parameters of surface layers by ultrasonic Love waves. 2016 IEEE International Ultrasonics Symposium Proceedings
  75. 75. Xi DM. Resonant second-harmonic generation accompanying nonlinear Love-wave propagation in an anisotropic waveguide. Ultrasonics Symposium of IEEE. 1997. pp. 578-580
  76. 76. Dorman J, Ewing M. Numerical inversion of seismic surface wave dispersion data and crust-mantle structure in the New York-Pennsylvania area. Geophysics. 67:5227-5241
  77. 77. Kiełczyński P, Szalewski M, Balcerzak A. Inverse procedure for simultaneous evaluation of viscosity and density of Newtonian liquids from dispersion curves of Love waves. Journal of Applied Physics. 2014;116:044902
  78. 78. Kalyanasundarm N. International Journal of Engineering Sciences. 1981;19:287
  79. 79. Frosch RA, Green PE Jr. The concept of a large aperture seismic array. Proceedings of the Royal Society A. 1966;290:368-384
  80. 80. Nakanishi I, Anderson DL. Measurements of mantle wave velocities and inversion for lateral heterogeneity and anisotropy analysis by the single-station method. Geophysical Journal Royal Astronomical Society. 1984;78:573-617
  81. 81. Haskell NA. Crustal reflection of plane SH waves. Journal of Geophysical Research. December 1960;65(12):4147-4150
  82. 82. Bataille K, Lund F. Nonlinear waves in elastic media. Physica. 1982/83;6D(1):95-104
  83. 83. Qu C. Piezoelectric vibration energy harvester actuated by seismic excitation [dissertation]. Politecnico di Milano; 2016
  84. 84. Penza M, Aversa P, Cassano G, Wlodarski W, Kalantarzadeh K, Layered SAW. Gas sensor with single-walled carbon nanotube-based nanocomposite coating. Sensors and Actuators B. 2007;127:168-178
  85. 85. Kiełczyński P, Szalewski M. An inverse method for determining the elastic properties of thin layers using Love surface waves. Inverse Problems in Science and Engineering. 2011;19:31-43
  86. 86. Guo FL, Wang GQ, Rogerson GA. Inverse determination of liquid viscosity by means of the Bleustein–Gulyaev wave. International Journal of Solids and Structures. 2012;49:2115-2120
  87. 87. Ueda K, Kondoh J. Estimation of liquid properties by inverse problems analysis based on shear horizontal surface acoustic wave sensor responses. Japanese Journal of Applied Physics. 2017;56:07 JD08
  88. 88. Chen X, Jiu Hui W, Kuan L, Cao P. Slow light and stored light of Love wave within a thin film. Optik. 2017;132:329-336
  89. 89. Zhang S, Bin G, Zhang H, Feng X-Q, Pan R, Alamusi NH. Propagation of Love waves with surface effects in an electrically-shorted piezoelectric nanofilm on a half-space elastic substrate. Ultrasonics. 2016;66:65-71

Written By

Piotr Kiełczyński

Submitted: 08 May 2017 Reviewed: 15 February 2018 Published: 10 April 2018