769 Photoacoustic 27. Photoacoustic Characterization of Materials The basic principles of photoacoustic generation of ultrasonic waves and applications to materials characterization of solid structures are discussed in this chapter. Photoacoustic techniques are a subset of ultrasonic methods wherein stress waves are used to obtain information about structural and material properties. In photoacoustic techniques, the ultrasound is typically generated using lasers, thereby enabling noncontact nondestructive characterization of the material properties of structures. Photoacoustic techniques have found application over a wide range of length scales ranging from macrostructures to nanometer-sized thin films and coatings. In this chapter, the basics of photoacoustics primarily as they relate to nondestructive characterization of solid materials are discussed. In Sect. 27.1, the basics of stress waves in solids is outlined. In Sect. 27.2, the process of photoacoustic generation is described. The major techniques of optical detection of ultrasound are then described in Sect. 27.3. The final section of this chapter is then devoted to some representative recent applications of photoacoustic characterization of materials. The objective here is to describe the basic principles involved, and to provide illustrative applications which take specific advantage of some of the unique features of the technique. Photoacoustics (also known as optoacoustics, laser ultrasonics, etc.) deals with the optical generation and detection of stress waves in a solid, liquid or gaseous medium. Typically, the technique uses modulated laser irradiation to generate high-frequency stress waves (ultrasonic waves) by either ablating the medium or through rapid thermal expansion. The resulting stress 27.1 Elastic Wave Propagation in Solids ......... 27.1.1 Plane Waves in Unbounded Media . 27.1.2 Elastic Waves on Surfaces .............. 27.1.3 Guided Elastic Waves in Layered Media ......................... 27.1.4 Material Parameters Characterizable Using Elastic Waves 770 771 772 774 776 27.2 Photoacoustic Generation ..................... 27.2.1 Photoacoustic Generation: Some Experimental Results............ 27.2.2 Photoacoustic Generation: Models . 27.2.3 Practical Considerations: Lasers for Photoacoustic Generation ........ 777 27.3 Optical Detection of Ultrasound ............. 27.3.1 Ultrasonic Modulation of Light ....... 27.3.2 Optical Interferometry................... 27.3.3 Practical Considerations: Systems for Optical Detection of Ultrasound . 783 783 785 777 780 783 789 27.4 Applications of Photoacoustics............... 789 27.4.1 Photoacoustic Methods for Nondestructive Imaging of Structures ................................ 789 27.4.2 Photoacoustic Methods for Materials Characterization ........ 793 27.5 Closing Remarks ................................... 798 References .................................................. 798 wave packets are also typically measured using optical probes. Photoacoustics therefore provides a noncontact way of carrying out ultrasonic interrogation of a medium to provide information about its properties. Photoacoustics can be used for nondestructive imaging of structures in order to reveal flaws in the structure, as well as to obtain the material properties of the structure. Part C 27 Sridhar Krishnaswamy 770 Part C Noncontact Methods Photoacoustic measurement systems are particularly attractive for nondestructive structural and materials characterization of solids because: • • Part C 27.1 • • • • they are noncontact; they can be nondestructive if the optical power is kept sufficiently small; they can be used for in situ measurements in an industrial setting; they are couplant independent (unlike contact acoustic techniques), providing absolute measurements of ultrasonic wave displacements; they have a very small footprint and so can be operated on curved complex surfaces; they are broadband systems providing information from the kHz to the GHz range, enabling the probing of macrostructures to very thin films. Over the past two decades, photoacoustic methods have evolved from being primarily laboratory research tools, which worked best on highly polished optically reflective specimens, to being incorporated in a wide range of industries for process monitoring applications. Photoacoustic methods are currently being used for imaging of flaws in real composite aerospace structures. They have been used for in situ process control in steel mills to measure the thickness of rolled sheets on the fly. Photoacoustic metrology tools are also widely used in the semiconductor industry for making wafer thickness measurements among other things. For a more comprehensive discussion of the principles of photoacoustic metrology, the reader is referred to the books by Scruby and Drain [27.1] and Gusev and Karabutov [27.2], as well as to several excellent review articles on laser generation of ultrasound [27.3], and optical detection of ultrasound [27.4–6]. 27.1 Elastic Wave Propagation in Solids There are many excellent books on stress waves in solids [27.7–9]. Here, only a selective review of elastic waves in solids is given. The field equations of linear elastodynamics are • the equations of motion ∇ · σ = ρü • σij, j = ρü i , (27.1) the constitutive relations σ = C·ε • → → σij = Cijkl εkl , (27.2) σ I = (σ11 σ22 σ33 σ23 =σ32 σ31 =σ13 σ12 =σ21 )T and the strain–displacement relations 1 ε= ∇u + (∇u)T 2 → view of these symmetries, there are at most 21 independent elastic stiffness constants for the most anisotropic material. With increasing levels of material symmetry, the number of independent elastic stiffness constants decreases, with only three for cubic crystals, and only two for isotropic materials. For simplicity, a contracted notation is often preferred. The six independent components of the stress and strain tensors are stacked up as six-dimensional column vectors: εI = (ε11 ε22 ε33 2ε23 =2ε32 2ε31 =2ε13 2ε12 =2ε21 )T . 1 εkl = (u k,l + u l,k ) , 2 (27.3) where σ is the stress tensor field, ε is the infinitesimal strain tensor field, C is the elasticity tensor, u is the displacement vector field, and ρ is the material density. Superposed dots imply time differentiation. From angular momentum balance considerations of nonpolar media, it can be shown that the stress tensor σ is symmetric, i.e., σij = σ ji . From its definition, the strain tensor ε is symmetric as well: εij = ε ji . It follows therefore, that the elasticity tensor has the following minor symmetries: Cijkl = C jikl and Cijkl = Cijlk . From thermodynamic considerations, the elasticity tensor also has the following major symmetry: Cijkl = Cklij . In (27.4) Thus, ij = 11 → I = 1; ij = 22 → I = 2; ij = 33 → I = 3; ij = 23 or 32 → I = 4; ij = 31 or 13 → I = 5; ij = 12 or 21 → I = 6. Capital subscripts will be used whenever the contracted notation is used. The constitutive relations in contracted notation then become ⎞⎛ ⎞ ⎛ ⎞ ⎛ σ1 c11 c12 c13 c14 c15 c16 ε1 ⎟⎜ ⎟ ⎜ ⎟ ⎜ ⎜σ2 ⎟ ⎜c12 c22 c23 c24 c25 c26 ⎟ ⎜ε2 ⎟ ⎟⎜ ⎟ ⎜ ⎟ ⎜ ⎜σ3 ⎟ ⎜c13 c23 c33 c34 c35 c36 ⎟ ⎜ε3 ⎟ ⎟⎜ ⎟ ⎜ ⎟=⎜ ⎜σ ⎟ ⎜c c c c c c ⎟ ⎜ε ⎟ ⎜ 4 ⎟ ⎜ 14 24 34 44 45 46 ⎟ ⎜ 4 ⎟ ⎟⎜ ⎟ ⎜ ⎟ ⎜ ⎝σ5 ⎠ ⎝c15 c25 c35 c45 c55 c56 ⎠ ⎝ε5 ⎠ σ6 c16 c26 c36 c46 c56 c66 ε6 (27.5) → σ I = C I J εJ . Photoacoustic Characterization of Materials As will be seen later in this chapter, photoacoustic measurements can be useful in determining the anisotropic elastic stiffness tensor. 27.1.1 Plane Waves in Unbounded Media u(r, t) = U exp[ik( · r − vt)] , (27.6) where U is the displacement amplitude vector, is a unit vector along the propagation direction of the wave, r = x1 ê1 + x2 ê2 + x3 ê3 is the position vector, k = 2π/λ is the wavenumber, λ is the wavelength, and v is the phase velocity of the wave. The angular frequency of the harmonic wave is related to the wavenumber and velocity through ω = kv = 2πv/λ. The êi are unit vectors√along the 1, 2, and 3 directions, and the symbol i = −1. Substituting the above into the field equations (27.1)–(27.3), results in the so-called Christoffel equation (27.7) Γik − δik ρv2 Ui = 0 , where Γik = Cijkl j l (27.8) is called the Christoffel matrix. The existence of plane waves propagating along any direction in a general anisotropic unbounded media follows directly from the existence of real solutions to the eigenvalue problem represented by (27.7). The eigenvalues are obtained by solving the secular equation det(Γik − δik ρv2 ) = 0 . nondispersive (i. e., the phase velocity is independent of the frequency). Again from the spectral theorem, corresponding to each of the real eigenvalues there is at least one real eigenvector and, furthermore, one can always find three orthogonal eigenvectors, say, U (i) . Therefore, in any homogeneous anisotropic material, one can always propagate three types of plane harmonic waves along any chosen propagation direction . In general, these three waves will have different phase velocities v(i) , and the corresponding particle displacement vectors U (i) will be mutually orthogonal. Each of these modes is called a normal mode of propagation. The direction of the displacement vector is called the polarization direction of the wave. Note, however, that the particle displacement vector need not be parallel or perpendicular to the propagation direction in general. If the polarization direction of a wave is parallel to the propagation direction, the wave is called a pure longitudinal wave. Waves with polarization direction normal to the propagation direction are pure shear waves. If the polarization directions are neither parallel nor perpendicular to the propagation direction, the waves are neither pure longitudinal nor pure shear. In such cases, the mode whose polarization makes the smallest angle to the propagation direction is called a quasi-longitudinal wave, and the other two are called quasi-shear waves. Plane Waves in Unbounded Isotropic Media For an isotropic material, as all directions are equivalent, consider a convenient propagation direction such as = ê1 so that the Christoffel matrix readily simplifies to ⎞ ⎛ c11 0 0 ⎟ ⎜ (27.10) = ⎝ 0 c14 0 ⎠ . 0 (27.9) It can be readily seen that the Christoffel matrix is symmetric, and under some nonrestrictive conditions on the elastic stiffness tensor, one can show that it is also positive definite. From the spectral theorem for positivedefinite symmetric matrices, it follows that there are three positive, real eigenvalues for Γij . This implies that the plane-wave phase velocities v (which are just the square root of these eigenvalues divided by the density) are guaranteed to be real and so will represent propagating modes. It should also be noted that, as the eigenvalues are independent of the frequency (and there are no boundary conditions to be satisfied here), the plane waves in an unbounded anisotropic media are 0 c14 The secular or characteristic equation then becomes c11 − ρv2 c44 − ρv2 c44 − ρv2 = 0 , (27.11) whose roots are v (1) ≡ vL = c11 ρ (27.12) with the corresponding polarization parallel to the propagation direction, which therefore represents a pure longitudinal wave; and two degenerate roots c44 (27.13) v(2) = v(3) ≡ vT = ρ 771 Part C 27.1 Consider a homogenous unbounded linear elastic anisotropic medium. Seek plane harmonic waves in such a medium given by the following displacement field: 27.1 Elastic Wave Propagation in Solids 772 Part C Noncontact Methods with the corresponding polarization along any direction on the plane perpendicular to the propagation direction, which therefore represent pure shear waves. Part C 27.1 Plane Waves in Unbounded Anisotropic Media For a general anisotropic material, it is not easy to simplify the secular equation analytically for arbitrary propagation directions, even though it may be possible to obtain analytically tractable expressions for special cases of propagation along certain material symmetry directions. In general, however, one seeks numerical solutions for the anisotropic problem. It is customary to represent the inverse of the phase velocity along any given propagation direction by means of slowness surfaces in so-called k-space with axes k i /ω (which is the reciprocal of velocity, hence slowness). Any direction in this space represents the propagation direction, and the distance to the slowness surface gives the reciprocal of the phase velocity of the associated mode in this direction. Figure 27.1 shows the slowness curves for cubic Si material. Note that for an isotropic material the slowness surfaces are just spheres, the inner one representing the longitudinal mode, and the two shear slowness surfaces being degenerate. Group Velocity The group velocity (which is the velocity with which a non-monochromatic wave packet of finite frequency content propagates in a general dispersive medium) is given by (g) Vj = ∂ω , ∂k j (27.14) where k j = k j and ω(k) in general. It can be shown that the group velocity (which is also the direction of energy flow) is always perpendicular to the slowness surface. It should be noted that it is usually the group velocity that is measured in experiments. 27.1.2 Elastic Waves on Surfaces Surface waves are waves that propagate along the surface of a body, and which typically decay in amplitude very rapidly perpendicular to the surface. Consider a half-space of an anisotropic but homogeneous medium (Fig. 27.2). In this case, in addition to the field equations of motion (27.1)–(27.3), the top surface (x3 = 0) is assumed to be traction free: σi3 = Ci3kl εkl = Ci3kl u k,l = 0 on x3 = 0 for i = 1, 2, 3 . (27.15) Seek so-called inhomogeneous plane wave solutions of the form u(r, t) = U exp[ik 3 x3 ] exp[ik( 1 x1 + 2 x2 − vt)] , (27.16) × 10– 4 2 kl2/ω [001] 1.5 1 Pure shear [010] ρ 1/2 Polarized: ⎛ ⎛ c ⎝ 44 ⎝ 0.5 [100] 0 kl1/ω –0.5 Quasi-longitudinal –1 Quasi-shear –1.5 –2 –2 –1.5 –1 – 0.5 0 Fig. 27.1 Slowness curves for silicon 0.5 1 1.5 2 × 10– 4 where it is required that the displacements decay with depth (x3 -direction) and the propagation vector be restricted to the x1 –x2 plane, i. e., = 1 ê1 + 2 ê2 . Such waves are called Rayleigh waves. Note that the 3 term is not to be thought of as the x3 component of the propagation vector (which is confined to be on the x1 –x2 plane), but rather a term that characterizes the decay of the wave amplitudes with depth. In fact, both 3 and the wave velocity v are to be determined from the solution to the boundary-value problem. If 3 = 0 or is pure real, then the wave does not decay with depth, and is not a Rayleigh wave. If 3 is complex, then in order to have finite displacements at x3 → ∞, we note that the imaginary part of 3 must be positive. This is the so-called radiation condition. Furthermore, to have a propagating wave mode, the velocity v must be positive and real. Here again the equations of motion obviously formally reduce to the same Christoffel equation (27.7), where the Christoffel matrix is again given by Γik = Cijkl j l . The solution to the eigenvalue problem Photoacoustic Characterization of Materials 3 x αn U (n) exp ik (n) 3 3 u(r, t) = n=1 × exp[ik( 1 x1 + 2 x2 − vt)] , (27.17) where αn are weighting constants. Using the above displacement field in the traction-free boundary conditions (27.15) results in a set of three homogenous equations for the weighting constants αn : σi3 =[ik] 3 x C3 jkl αn Uk(n) l(n) exp ik (n) 3 3 Swapping out the elastic stiffness in favor of the longitudinal and transverse velocities vL and vT one obtains 2 2 2 vT 3 + vT2 − v2 vL2 23 + vL2 − v2 = 0 , (27.22) whose six roots are ⎧ 2 1/2 ⎪ ⎪ v ⎪ ±i 1 − , ⎪ vT ⎪ ⎪ ⎪ ⎨ 2 1/2 3 = ±i 1 − vv , T ⎪ ⎪ ⎪ 1/2 ⎪ ⎪ 2 ⎪ ⎪ . ⎩±i 1 − vvL =0 , (27.18) where (n) 1:2 = 1:2 is used for simplicity of notation. The above can be cast as dmn αn = 0 , (27.19) where the elements of the 3 × 3 d matrix are: dmn = C3mkl Uk(n) l(n) (no sum over (n) intended here). Nontrivial solutions to the above are obtained if det(dmn ) = 0 . (27.20) Rayleigh Waves on Isotropic Media The isotropic surface wave problem is analytically tractable. Since all directions are equivalent, pick = ê1 as the propagation direction. The corresponding secular equation is 2 1 1 (C11 − C12 ) 23 + (C11 − C12 ) − ρv2 2 2 × C11 23 + C11 − ρv2 = 0 . (27.21) (27.23) Since it is required that the displacements decay with depth, and since the velocities have to be real, the three admissible roots for the Rayleigh wave velocity v ≤ vT < vL are only those corresponding to the negative exponential. The corresponding eigenvectors are ⎛ ⎞ 0 ⎜ ⎟ U (1) = ⎝1⎠ , 0 ⎛ ⎞ 2 1/2 vT v i 1 − ⎜ v ⎟ vT ⎜ ⎟ U (2) = ⎜ ⎟, ⎝ ⎠ 0 vT v vT v ⎛ n=1 × exp[ik( 1 x1 + 2 x2 − vt)] 773 ⎞ ⎜ ⎟ ⎜ ⎟ 0 U (3) = ⎜ 2 1/2 ⎟ ⎝ v ⎠ −i vL 1 − vvL (27.24) and the boundary conditions (27.19) become ⎛ ⎞ 0 (U3(2)+U1(2) (2) (U3(3)+U1(3) (3) 3 ) 3 ) ⎜ (1) ⎟ ⎝ 3 ⎠ 0 0 (2) (2) (2) (3) (3) (3) 0 (c11 U3 3 +c12 U1 ) (c11 U3 3 +c12 U1 ) ⎛ ⎞ ⎛ ⎞ α1 0 ⎜ ⎟ ⎜ ⎟ × ⎝α2 ⎠ = ⎝0⎠ . (27.25) α3 0 The determinant of the matrix above vanishes for two cases. One case corresponds to a shear wave in the bulk of the material with no decay with depth and this is not a surface wave. The other case leads to the characteristic Rayleigh wave equation vT2 3 =0, β − 8(β − 1) β − 2 1 − 2 vL β= v2 . vT2 (27.26) Part C 27.1 again formally yields the same secular equation (27.9), except that in this case this leads to a sixth-order polynomial equation in both 3 and v, both of which are as yet undetermined. It is best to think of this as a sixth-order polynomial equation in 3 with velocity as a parameter. In general, there will be six roots for 3 (three pairs of complex conjugates). Of these, only three are admissible so that the waves decay with depth according to the radiation condition. Denote the admissible values by (n) 3 , n = 1, 2, 3 and the corresponding eigenvectors as U (n) . That is, there are three possible surface plane harmonic wave solutions, and the most general solution for the displacement field is a linear combination of these three 27.1 Elastic Wave Propagation in Solids 774 Part C Noncontact Methods This can be solved numerically for any given isotropic material. Let the velocity corresponding to the solution to (27.26) be called vR . The corresponding Rayleigh wave displacement field is given by Part C 27.1 u(r, t) = U0 exp[ik(x 1 − vR t)] ⎡ ⎤ ⎤ ⎛ ⎡ 1 1 1 1 ⎞ 2 2 2 4 2 4 2 2 vR vR vR vR ⎣ ⎣ ⎦ ⎜exp k 1− 2 x 3 − 1− 2 1− 2 exp k 1− 2 x 3⎦⎟ ⎜ ⎟ vL vL vL vT ⎜ ⎟ ⎜ ⎟ 0 ⎜ ⎟ ⎧ ⎫ ⎤ ⎡ ⎜ ⎟ 1 ⎪ ⎪ ⎟ 2 2 ⎪ ⎪ ×⎜ v ⎥ ⎪ ⎟, ⎢ ⎪ ⎤ ⎜ ⎪ ⎡ x 3⎦ ⎪ exp⎣k 1− R 1⎪ 1 ⎪ ⎜ ⎟ 2 ⎨ ⎬ vT 2 2 ⎜ ⎟ 2 2 ⎣k 1− vR x 3⎦ − ⎜ −i 1− vR ⎟ exp 2 2 ⎜ 1 1 ⎪ ⎟ vL v ⎪ 2 4 2 4 ⎪ ⎠ ⎪ L ⎝ vR vR ⎪ ⎪ ⎪ ⎪ 1− 1− ⎪ ⎪ ⎩ ⎭ v2 v2 L L (27.27) where U0 is the amplitude. It should be noted that, for Rayleigh waves in isotropic media, the vectorial displacement is entirely in the plane contained by the propagation vector and the normal to the surface (the so-called sagittal plane). The x3 decay of the two nonzero components is not equal. The u 3 component is phase delayed with respect to the u 1 component by 90◦ , and they are not of equal magnitude. Therefore, the particle displacement is elliptical in the sagittal plane. Also note that the velocity is independent of the frequency, and therefore the waves are nondispersive. may be larger than the slow (quasi-)transverse velocity for the bulk material. Along neighboring directions to these isolated directions, a pseudo-Rayleigh wave solution exists which actually does not quite satisfy the displacement decay condition at infinite depth. (These pseudo-Rayleigh waves have energy vectors with nonzero component perpendicular to the surface.) In practice, these waves can actually be observed experimentally if the source-to-receiver distance is not too great. 27.1.3 Guided Elastic Waves in Layered Media Elastic waves can also be guided in structures of finite geometry. Guided waves in rods, beams, etc. are described in the literature. Of particular importance is the case of guided waves in layered media such as composite materials and coatings [27.11]. The simplest case is that of guided waves in plates. Consider an infinitely wide plate of thickness h made of an anisotropic but homogeneous medium (Fig. 27.3). Here, in addition to the field equations of elastodynamics, there are two traction-free boundary condition to be satisfied on the two end planes x3 = 0, h. Once again seek solutions of the form u(r, t) = U exp[ik 3 x3 ] exp[ik( 1 x1 + 2 x2 − vt)] , (27.28) Rayleigh Waves on Anisotropic Crystals For a general anisotropic medium, the solution to the Rayleigh wave problem is quite involved and is typically only obtainable numerically. It has been shown [27.10] that Rayleigh waves exist for every direction in a general anisotropic medium, but their velocity depends on the propagation direction. In some cases, the decay term 3 can be complex and not just pure imaginary, which means that the decay with x3 can be damped oscillatory. Along certain isolated directions, the Rayleigh surface wave phase velocity which represents an inhomogeneous plane wave propagating parallel to the plane of the plate with an amplitude variation along the thickness direction given by the 3 term. Again, as for the surface wave case, 3 is at this point undetermined along with the wave velocity v, but the depth decay condition is no longer valid. As in the Rayleigh wave problem, solving the Christoffel equation will lead to six roots for 3 (three pairs of complex conjugates), but now all of these are admissible. Let the eigenvalues be denoted by (n) 3 , n = 1, 2, . . . , 6, x1 x2 x1 x2 x3 l x3 h Fig. 27.2 Surface waves on a half-space Fig. 27.3 Lamb waves in thin plates Photoacoustic Characterization of Materials and let U (n) be the corresponding eigenvectors. There are therefore six plane-wave solutions to the Christoffel equation, each of which is called a partial wave. The displacement field is therefore a linear combination of the six partial wave solutions Ωsym (ω,v,vL ,vT ) = (k2−β 2 )2 cos(αh/2) sin(βh/2) + 4k2 αβ sin(αh/2) cos(βh/2) =0 , (27.29) where σi3 = (ik) 6 Ci3kl αn Uk(n) l(n) exp −ik (n) 3 x3 n=1 × exp [ik ( 1 x1 + 2 x2 − vt)] . (27.30) Using the above in the traction-free boundary conditions leads to a set of six homogenous equations: dα = 0 , (27.31) where the d matrix is now 6 × 6 and is given by dmn (1,2,3) dmn = (4,5,6) dmn (n) = C3mkl Uk(n) l(n) eik 3 h , The first of the two dispersion relations corresponds to symmetric modes, and the second gives rise to antisymmetric modes. These dispersion relations are quite problematic because α, β can be real or complex depending on whether the phase velocity is less than or greater than the bulk longitudinal or bulk shear wave speeds in the material. Numerical solutions to the dispersion relations indicate that Lamb waves are highly dispersive and multimodal. Figure 27.4 shows Lamb wave dispersion for an isotropic aluminum plate. 12 000 10 000 (27.32) 8000 where for convenience of notation we define (n) 1:2 = 1:2 . Nontrivial solutions for the weighting constants are obtained if det(dmn ) = 0 . 1/2 v2 α = ik 1 − 2 , vL 1/2 v2 β = ik 1 − 2 . vT Lamb wave phase velocity (m/s) (1,2,3) = C3mkl Uk(n) l(n) dmn (no sum over n intended here) (4,5,6) dmn (27.33) This provides the dispersion relation for the guided waves. Note that, unlike for bulk waves and surface waves, now the solution depends on the wavenumber k. Therefore, we can now expect dispersive solutions where the velocity will depend on frequency. Guided Waves in Isotropic Plates For an isotropic plate, one can show that the above leads to so-called Lamb waves. Lamb wave dispersion parti- 6000 S0 4000 2000 0 A0 0 2 4 6 8 10 f h (MHz mm) Fig. 27.4 Lamb-wave dispersion curve for an aluminum plate. The horizontal axis is the frequency–thickness product and the vertical axis is the phase velocity. The fundamental antisymmetric mode A0 , the fundamental symmetric mode S0 , and higher-order modes are shown Part C 27.1 + 4k2 αβ cos(αh/2) sin(βh/2) (27.34) =0 , n=1 where αn are weighting constants. The corresponding stress components σi3 can then be readily expressed as 775 tions into two separate dispersion relations [27.7] Ωasym (ω,v,vL ,vT ) = (k2−β 2 )2 sin(αh/2) cos(βh/2) 6 αn U (n) exp ik (n) u(r, t) = 3 x3 × exp [ik ( 1 x1 + 2 x2 − vt)] , 27.1 Elastic Wave Propagation in Solids 776 Part C Noncontact Methods Part C 27.1 Guided Waves in Multilayered Structures Consider the N-layer structure on a substrate shown in Figure 27.5. Acoustic waves may be coupled into the structure either via air or a coupling liquid on the top surface. We are interested in the guided plane acoustic waves that can be supported in such a system with propagation in the plane of the structure. The six partial wave solutions obtained in Sect. 27.1.3 are valid for a single layer, and these can be used to assemble the solution to the multilayer problem. The solution approach is to determine the partial wave solutions for each layer, and thereby the general solution in each layer as a linear combination of these partial wave solutions. We then impose traction and displacement boundary and continuity conditions at the interfaces. This is most effectively done by means of the transfer matrix formulation [27.12, 13]. For this purpose, a state vector is created consisting of the three displacements and the three tractions that are necessary to ensure continuity conditions: S = (u 1 u 2 u 3 σ33 σ32 σ31 )T . (27.35) This vector must be continuous across each interface. For any single layer, following (27.29, 30), we have the state vector in terms of the weighting constants as follows: S = D(x3 ) (α1 α2 α3 α4 α5 α6 ) , T (27.36) where the common term exp[ik( 1 x1 + 2 x2 − vt)] is omitted for simplicity, and the matrix D(x3 ) is determined either analytically or numerically given the material properties and frequency and wavenumber. Note that the above state vector can be used for a substrate (treated as a half-space) by recognizing that in the R (θ) θ Coupling water/air Top Layer 1 h1 h2 Layer 2 .. . Plane wave hn Layer N Substrate: half-space x3 Fig. 27.5 Waves in layered structures T (θ) x1 Phase velocity (m/s) 8000 6000 Longitudinal velocity of Ti 4000 Shear velocity of Ti 2000 0 2 4 6 8 10 f h (MHz mm) Fig. 27.6 Guided wave dispersion curves for an aluminum layer on a titanium half-space (Al/Ti) substrate layer the radiation condition is valid, which means that for this last layer only three partial waves are admissible. For a single layer of finite thickness, the state vector at the two sides are related through direct elimination of the weighting constants vector: [S(−) (x3 = 0)] = [D(x3 = 0)][D(x3 = h)]−1 [S(+) (x3 = h)] ≡ [T1 ][S(+) (x3 = h)] , (27.37) where T1 is called the transfer matrix of the layer. For an N-layer system, it readily follows from continuity conditions that the state vector at the top and bottom surfaces are related through [S(−) ] = [T1 ][T2 ][T3 ] . . . [T N ][S(+) ] ≡ [T][S(+) ] , (27.38) where T is called the global transfer matrix of the N-layer system. Given the boundary conditions for the top and bottom surfaces, or the radiation condition if the bottom layer is a substrate (half-space), the above provides the appropriate dispersion relations for the various guided-wave modes possible. This can be solved numerically for a given system. For illustrative purposes, Fig. 27.6 shows the guided-wave dispersion curves for an aluminum layer on a titanium substrate. 27.1.4 Material Parameters Characterizable Using Elastic Waves Since elastic wave propagation in a solid depends on the material properties and the geometry of the struc- Photoacoustic Characterization of Materials measurements can often be made independent of the support conditions of the structure. An example of this will be seen in section Sect. 27.4.2 on photoacoustic characterization of ultrathin films. It should be noted that techniques based on interrogation using linear elastic waves can only access stiffness properties and cannot in general reveal information about the strength of a material. If large-amplitude stress waves can be generated, material nonlinearity can be probed. However, we will not address this issue here since photoacoustic generation is generally kept in the so-called thermoelastic regime where the stress waves are typically of small amplitude. 27.2 Photoacoustic Generation Photoacoustics arguably traces its history back to Alexander Graham Bell, but the field really got its impetus in the early 1960s starting with the demonstration of laser generation of ultrasound by White [27.14]. Since then, lasers have been used to generate ultrasound in solids, liquids, and gases for a number of applications. A comprehensive review of laser generation of ultrasound is given in Hutchins [27.3] and Scruby and Drain [27.1]. Here we will restrict attention to the generation of ultrasound in solids using pulsed lasers. The basic mechanisms involved in laser generation of ultrasound in a solid are easy to outline. A pulsed laser beam impinges on a material and is partially or entirely absorbed by it. The optical power that is absorbed by the material is converted to heat, leading to rapid localized temperature increase. This results in rapid thermal expansion of a local region, which in turn leads to generation of ultrasound into the medium. If the optical power is kept low enough that the material does not melt and ablate, the generation regime is called thermoelastic (Fig. 27.7a). If the optical power is high enough to lead to melting of the material and plasma formation, once again ultrasound is generated, but in this case via momentum transfer due to material ejection (Fig. 27.7b). The ablative regime of generation is typically not acceptable for nondestructive characterization of materials, but is useful in some process monitoring applications especially since it produces strong bulk wave generation normal to the surface. In some cases where a strong ultrasonic signal is needed but ablation is unacceptable, a sacrificial layer (typically a coating or a fluid) is used either unconstrained on the surface of the test medium or constrained between the medium and an optically transparent plate. The sacrificial layer is then ablated by the laser, again leading to strong ultrasound generation in the medium due to momentum transfer. 27.2.1 Photoacoustic Generation: Some Experimental Results Before delving into the theory of photoacoustic generation, it is illustrative to look at some of the kinds of ultrasonic signals that can be generated using photoacoustic generation. Photoacoustic Longitudinal and Shear Wave Generation Using a Point-Focused Laser Source Figure 27.8 shows the surface normal displacement at the epicentral location generated by a point-focused thermoelastic source and measured using a homodyne interferometer (described in Sect. 27.3.2). The arrival a) b) Thermoelastic expansion Thermoelastic expansion and momentum transfer due to material ejection x3 x1 Ablation Generating laser Generating laser Fig. 27.7a,b Laser generation of ultrasound in (a) the thermoelastic regime and (b) the ablative regime 777 Part C 27.2 ture, elastic waves can be used to measure some of these parameters. For instance, by measuring the bulk wave speeds of an anisotropic medium in different directions, it is possible to obtain the elasticity tensor knowing the density of the material. Using guided acoustic waves in multilayered structures, it may be possible to infer both material stiffnesses as well as layer thicknesses. The primary advantage of using elastic waves over performing standard load–deflection type of measurements arises from the fact that these measurements can be made locally. That is, while load–deflection measurements require accurate knowledge of the boundary or support conditions of the entire structure, elastic wave 27.2 Photoacoustic Generation 778 Part C Noncontact Methods 0 Epicentral displacement –8 –10 Part C 27.2 –12 –14 Thermoelastic generation Longitudinal wave arrival –16 0.8 1 1.2 1.4 1.6 1.8 2 2.2 Time (μs) Fig. 27.8 Experimentally measured epicentral surface normal displacement of a longitudinal wave is seen distinctly, followed by a slowly increasing wash, until the shear wave arrives. a) The directivity (i. e., the wave amplitude at different angles to the surface normal) of the longitudinal and shear waves generated by a thermoelastic source have been calculated theoretically and measured experimentally by a number of researchers [27.1]. The longitudinal and shear wave directivity using a shear dipole model (described later) for the thermoelastic laser source have been obtained using a mass–spring lattice elastodynamic calculation [27.15] and are shown in Fig. 27.9. It is seen that, for thermoelastic generation, longitudinal waves are most efficiently generated in directions that are at an angle to the surface normal. Epicentral longitudinal waves are much weaker. This is consistent with the results of other models as well as with those from experiments [27.16]. Photoacoustic Rayleigh Wave Generation Using a Point- and Line-Focused Laser Source Thermoelastic generation of Rayleigh waves has been extensively studied (see Scruby and Drain [27.1]). a) Displacement (arb. units) 90 120 60 150 30 180 0 b) Displacement (arb. units) Laser source b) Rayleigh wave 90 120 60 150 30 180 0 Laser source Fig. 27.9a,b Calculated directivity pattern of thermoelastic laser generated ultrasound: (a) longitudinal waves and (b) shear waves 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 Time (μs) Fig. 27.10a,b Rayleigh wave generation using a pointfocused laser source. (a) Experimental results for surface normal displacements. (b) Theoretical calculations of the surface normal (solid) and horizontal (dotted) displacements Photoacoustic Characterization of Materials 27.2 Photoacoustic Generation 779 Displacement (arb. units) a) Displacement (arb. units) 5 0 0 5 10 15 20 25 –5 –15 Rayleigh wave 0 2 4 6 8 Time (μs) Photoacoustic Guided-Wave Generation in Multilayered Structures Thermoelastic generation of ultrasound on multilayered structures has also been demonstrated, motivated by potential applications to the characterization of coatings. In these cases, again the various guided modes are gen- b) Displacement (arb. units) 0.1 0 –0.1 Rayleigh wave –0.2 –0.3 0 2 a thin aluminum plate. The surface normal displacements at some distance from the generating source shows the dispersive nature of Lamb waves 4 a) Displacement (arb. units) 6 8 Time (μs) Fig. 27.11a,b Surface normal displacement showing a monopolar Rayleigh wave from a thermoelastic line source: (a) theoretical and (b) experimental Shown in Fig. 27.10a are measurements of the far-field surface normal displacement made using a homodyne interferometer (discussed in Sect. 27.3.2). Figure 27.10b shows the theoretically calculated surface displacements in the far field of a thermoelastic point source, using a shear-dipole model [27.15]. A strong bipolar Rayleigh wave is seen to be generated by the thermoelastic source. If the laser source is a line source rather than a point-focused source, the Rayleigh wave becomes a monopolar pulse, as shown both theoretically and experimentally in Fig. 27.11 [27.15]. Photoacoustic Lamb Wave Generation Thermoelastic Lamb wave generation in thin plates has been experimentally studied by Dewhurst et al. [27.17], and has been modeled by Spicer et al. [27.18]. Figure 27.12 shows the Lamb waves that are generated by a single thermoelastic line source on a thin aluminum plate. The detection system used was a broadband twowave mixing interferometer (Sect. 27.3.2). The figure shows the dispersive nature of Lamb wave propagation. 0 200 400 600 800 Time (ns) 600 800 Time (ns) b) Displacement (arb. units) 0 200 400 Fig. 27.13a,b Thermoelastic generation on a 2.2 μm Ti thin-film coating on an aluminum substrate. Surface normal displacement at a distance of 1.5 mm (a) measured experimentally and (b) predicted by the theoretical model Part C 27.2 Fig. 27.12 Thermoelastically generated Lamb waves in –10 –20 30 Time (μs) 780 Part C Noncontact Methods Part C 27.2 erally dispersive. Murray et al. [27.19] have performed comparisons of theoretical and experimental signals in layered structures on a substrate. High-frequency ultrasonic guided waves were generated with a miniature Nd:YAG laser with a 1 ns pulse width and 3 μJ per pulse, focused down to a point source of about 40 μm diameter. Figure 27.13a shows the ultrasonic surface displacements measured using a broadband stabilized Michelson interferometer (described in Sect. 27.3.2) at a distance of 1.5 mm from the laser source. The specimen was a 2.2 μm Ti coating on aluminum. Also shown in Fig. 27.13b are the theoretically expected signals using a thermoelastic model for the two-layer case. It is seen that the thermoelastic model adequately captures the experimental behavior. From these examples, it is clear that photoacoustic generation of bulk and guided waves is technically feasible using pulsed laser sources, and theoretical models for the process are well established. In the next section, the basic theoretical models for photoacoustic generation are outlined. 27.2.2 Photoacoustic Generation: Models It is important to characterize the ultrasound generated by laser heating of a material in order to determine the amplitude, frequency content, and directivity of the ultrasound generated. In general, models are also useful in extracting material property information from the measured data. There are two modes of photoacoustic generation in solids: ablative and thermoelastic. If the material ablates, the ultrasound that results from momentum transfer can be modeled as arising from a normal impulsive force applied to the surface. Analytical solutions to this problem can be obtained by appropriate temporal and spatial convolution of the elastodynamic solution for a point impulsive force on a half-space [27.7]. Recently, a complete model of ultrasound generation in the ablative regime has been given by Murray et al. [27.20]. As ablative generation is typically not used in nondestructive evaluation, we will not explore this further here. We will only consider thermoelastic generation. The basic problem of thermoelastic generation of ultrasound can be decomposed into three subproblems: (i) electromagnetic energy absorption by the medium (ii) the consequent thermal diffusion problem with heat sources (due to the electromagnetic energy absorbed) (iii) the resulting elastodynamic problem with volumetric sources (due to thermal expansion) There are two important cases that are relevant to materials characterization: (i) bulk-wave photoacoustics where typically onedimensional bulk waves of GHz frequency range are launched using femtosecond pulsed laser sources to characterize thin films and coatings (ii) guided-wave photoacoustics where nanosecond focused laser sources are used to launch kHz to MHz ultrasonic waves Guided-Wave Photoacoustic Generation For simplicity a fully decoupled linear analysis for homogeneous, isotropic materials is considered. The optical energy that is absorbed depends on the wavelength of the laser light and the properties of the absorbing material. The optical intensity variation with depth inside an absorbing medium that is illuminated by a light beam at normal incidence is given by an exponential decay relation I (x1 , x2 , x3 , t) = I0 (x1 , x2 , t) exp(−γ x3 ) , (27.39) where I0 (x1 , x2 , t)is the incident intensity distribution at the surface (which is a function of the laser parameters) and γ is an absorption coefficient characteristic of the material for the given wavelength of light. The optical energy absorbed by the material leads to a distributed heat source in the material given by q(x1 , x2 , x3 , t) = q0 (x1 , x2 , t)γ exp(−γ x3 ) , (27.40) where q0 is proportional to I0 and has the same spatial and temporal characteristics as the incident laser source. The corresponding thermal problem is then solved for the given thermal source distribution using the equations of heat conduction ∂T (27.41) = κ∇ 2 T + q , ρC ∂t where T is the temperature, ρ is the material density, C is the heat capacity at constant volume, and κ is the thermal diffusivity. The necessary thermal boundary conditions arise from the fact that there is no heat flux across the surface, and initially the medium is at uniform temperature. The temperature distribution can be calculated for a given laser source and material by solving the heat conduction equation (27.41). For most metals, heat diffusion can be significant and needs to be taken into Photoacoustic Characterization of Materials u = ∇φ + ∇ × ψ , (27.42) where φ and ψ are the scalar and vector potentials, respectively. The equations of motion including the volumetric expansion source then become 1 φ̈ = φT , vL2 1 ∇ 2 ψ − 2 ψ̈ = 0 , vT ∇2φ − (27.43) where vL and vT are the longitudinal and shear wave speeds of the material, respectively, and a superposed dot indicates time differentiation. The temperature rise from the laser energy absorbed leads to a volumetric expansion source given by φT = 3λ + 2μ αT T , λ + 2μ (27.44) where αT is the coefficient of linear thermal expansion, and λ and μ are the Lamé elastic constants of the material. The above wave equations need to be solved along with the boundary conditions that the surface tractions vanish. Given the laser source parameters, the resulting temperature and elastodynamic fields are obtained from the above system of equations using transform techniques [27.21, 22]. The heat conduction and the elastodynamic equations are transformed using onesided Laplace transform in time, and either a Fourier (for a line source) or Hankel (for a point source) transform in space. Closed-form solutions can be obtained analytically in the transformed domain, and numerically inverted back into the physical domain [27.21–23]. For most metallic materials, the absorption coefficient is high enough that the optical energy does not penetrate very much into the material. The optical penetration depth defined as 1/γ is on the order of a few nanometers for most metals over the optical wavelengths typically used for laser generation of ultrasound. If, in addition, the time scale of interest is such that significant thermal diffusion does not occur, and therefore the volumetric thermoelastic expansion source is confined to the surface region, it has been shown that the volumetric sources can be replaced by an equivalent traction boundary condition on the surface. Scruby et al. [27.24] have argued that the relevant elastodynamic problem is that of shear dipoles acting on the surface of the body. Their argument was based on the consideration that a point expansion source in the interior of a solid can be modeled as three mutually orthogonal dipoles [27.7], and this degenerates into a pair of orthogonal dipoles as the expansion source moves to a free surface (Fig. 27.7a). This approach was given a rigorous basis in the form of the surface center of expansion (SCOE) model proposed by Rose [27.25], which predicts all the major features that have been observed in thermoelastic generation. Specifically, if the optical penetration depth is very small (i. e., γ → ∞), and the laser beam is assumed to be focused into an infinitely long line along the x1 -direction, and with a delta-function temporal dependence, the resulting heat source simplifies to q(x1 , x2 , x3 , t) = Q 0 δ(x1 )δ(x3 )δ(t) , (27.45) where Q 0 is the strength of the heat source proportional to the laser energy input. If thermal diffusion is also neglected, the resulting simplified problem can be explicitly solved [27.21, 23]. The corresponding in-plane stresses are given by [27.23] σ31 = Dδ (x1 )H(t) , (27.46) where a prime indicates differentiation with respect to the argument, and H(t) is the Heaviside step function. The above indicates that the shear dipole model of Scruby et al. [27.24] is indeed valid in this limit of no thermal diffusion and no optical penetration. The dipole magnitude D is given by [27.23] D= αT Q 0 2μ (3λ + 2μ) . λ + 2μ ρC (27.47) Solutions to the more general case where the temporal and spatial characteristics are more typical of real laser pulses can be readily obtained from the above solution using a convolution over space and time. If optical penetration depth is significant or if thermal diffusion is important, solutions to the complete system of equations will have to be obtained numerically [27.22, 23]. The shear-dipole model can also be extended to multilayered structures by using the transfer matrix formulation described in Sect. 27.1.3. Murray et al. [27.19] have shown results for guided-wave photoacoustic generation in a two-layer system consisting of a 2.2 μm Ti 781 Part C 27.2 account by solving the full heat conduction equation. For insulators, heat conduction may be neglected and the resulting adiabatic temperature rise is readily obtained by setting κ → 0 in (27.41). Next the elastodynamic problem is considered. For isotropic materials, the equations of elastodynamics can be cast in terms of the scalar and vector displacement potentials [27.7]. The elastic displacement field u can be decomposed into 27.2 Photoacoustic Generation 782 Part C Noncontact Methods coating on an aluminum substrate. Typical results are shown in Fig. 27.13b. Part C 27.2 Bulk-Wave Photoacoustic Generation For photoacoustic characterization of very thin films and coatings, an alternate approach is to use very highfrequency (very short-wavelength) bulk longitudinal waves that are launched into the depth of the specimen. This technique has come to be known as picosecond ultrasonics [27.26]. This necessitates the use of unfocused femtosecond laser sources that impinge on the specimen surface. If the diameter of the heating region (typically tens of μm) is much larger than the film thickness (typically nanometers), and the optical skin depth is much less than the film thickness, a one-dimensional thermoelastic model can be used [27.27, 28]. The temperature field T therefore only varies with the depth direction x, and the only nonzero displacement component is the one along the x-direction, denoted by u. To solve the complete thermoelastic problem in multilayer structures, a photothermoelastic transfer matrix approach is adopted, as shown in Fig. 27.14. Following Miklos et al. [27.29] we apply the classical partially coupled thermoelastic differential equations for the heat transfer and wave propagation in homogeneous isotropic or cubic media as follows: κ ∂T ∂2 T − ρC = −q 2 ∂t ∂x ∂2u ∂2u ∂T c 2 −ρ 2 = λ , ∂x ∂x ∂t be expressed as λ = cαT , where αT is the thermal expansion coefficient, and q is the absorbed energy due to laser irradiation (27.40). The linear coupled thermoelastic equations (27.48) hold for every thin layer in the multilayer structure, with total thickness h, as shown in Fig. 27.14. The boundary conditions are assumed to be no heat transfer or tractions across the surfaces at x = 0 and x = h. It is also assumed that all thin layers are in perfect contact, which ensures continuity of displacement, temperature, heat flux J = −κ∂T /∂x, and elastic stress σ = c∂u/∂x − λT at every interface. The coupled problem given by (27.48) can be readily solved using the transfer-matrix formulation a) Reflection coefficient change (arb. units) 1.2 1 0.6 where ρ is the density, κ is the thermal conductivity, C is the specific heat per unit volume, c is the effective elastic stiffness, λ is the thermal stress tensor, which can Supported film 0.4 Unsupported film 0.2 0 (27.48) Acoustic echoes 0.8 0 100 200 300 400 Delay time (ps) b) Elastic strain (10– 4) Al 2 10 ps 30 ps 50 ps 70 ps 1 y Pump laser Layer 1 Layer 2 ··· Layer n-1 Layer n 0 Si3N4 –1 Heat flux and elastic wave d1 0 d2 x1 0 dn–1 x2 dn xn–1 xn (h) x Fig. 27.14 Geometry of the photothermoelastic model for multilayer structures 100 200 300 400 500 600 Depth (nm) Fig. 27.15 (a) Simulated transient optical reflectivity change due to femtosecond laser heating for both supported and unsupported 300 nm aluminum and 300 nm silicon nitride double-layer thin films; (b) corresponding thermoelastic strain pulse shape and propagation in the unsupported double-layer thin film Photoacoustic Characterization of Materials properties of the materials determine the thermoreflectivity signal. 27.2.3 Practical Considerations: Lasers for Photoacoustic Generation Several different types of lasers are commercially available. The material in which the ultrasound is to be generated, and the desired frequency content of the ultrasound, dictate the type of laser to use. The generating laser wavelength, its energy, its pulse duration, and its repetition rate are all parameters that can be selected based on applications. The repetition rate is important primarily for speed of testing. The pulse duration along with other parameters such as the spatial extent of the generating volume/area dictate the frequency content of the ultrasound generated. A modulated continuous-wave laser is possibly adequate for low-frequency (order of tens of kHz) generation. Typically, laser pulses on the order of 10 ns are used for the generation of ultrasound in the 100 kHz–10 Mhz frequency range. For even higher frequencies, pulse widths on the order of 100 ps (resulting in ultrasound in the order of 100 MHz frequency range) or even femtosecond (for GHz range) laser systems may be necessary. The optical power required depends on the material to be tested and whether laser damage is acceptable or not. Lasers with optical power ranging from nanojoules to microjoules to several hundred millijoules have all been used to produce ultrasound in structures. Finally, the choice of laser wavelength primarily depends on the material absorption. Laser wavelengths ranging from the ultraviolet to the infrared and higher have been used for laser generation of ultrasound. 27.3 Optical Detection of Ultrasound Optical detection of ultrasound is attractive because it is noncontact, with high detection bandwidth (unlike resonant piezoelectric transducers), and it can provide absolute measurement of the ultrasonic signal. In this section, the ways in which ultrasonic signal information can be encoded onto a light beam are first described, followed by a discussion of two methods of demodulating the encoded information using optical interferometry. For a more complete review of optical detection of ultrasound, the reader is referred to a number of excellent review articles on the subject [27.4–6]. 783 27.3.1 Ultrasonic Modulation of Light To monitor ultrasound optically, a light beam should be made to interact with the object undergoing such motion. Interaction of acoustic waves and light waves in transparent media has a long history (see, for instance, [27.32]), and will not be reviewed here. Attention will be confined here to opaque solids that either reflect or scatter light. In this case, the light beam can only be used to monitor the surface motion associated with the ultrasound. Part C 27.3 as described in Sect. 27.1.3 but now in the Laplacetransformed (over time) domain. As an illustrative example, bulk-wave photoacoustic generation in a 300 nm aluminum and 300 nm silicon nitride double-layer film with and without silicon substrate are shown in Fig. 27.15. Figure 27.15a shows the simulation of the transient thermoelastic strain that propagates in the unsupported film. The laser pulse duration used in the calculation is a 100 fs Hanning function [27.29]. The elastic strain generated on the near surface propagates toward the interface with the velocity of longitudinal waves in aluminum and arrives at the Al and Si3 N4 interface at about 50 ps. Part of the energy is transmitted into the silicon nitride layer and the rest is reflected back toward the front surface, as shown in Fig. 27.15a. Also shown in Fig. 27.15b is the transient change in the surface reflectivity due to temperature and strain changes induced by the laser. It should be pointed out that, in picosecond ultrasonics, the transient reflectivity is typically monitored in a snapshot manner using a femtosecond laser pulse that is progressively time delayed with respect to the photoacoustic generation pulse (see [27.30, 31], for instance). This is necessitated by the very high temporal resolutions needed to monitor the transient photothermoacoustic phenomena that arise from femtosecond laser irradiation. The initial transient rise and subsequently exponential-like decay process are due to the thermoreflectivity changes. The superimposed multiple spikes are due to the acoustic waves that are reflected from the Al/Si3 N4 interface and the Si3 N4 /silicon or air interface. The elastic properties and density of each layer affect the acoustic wave arrival time as well as the signal amplitude, while the thermal 27.3 Optical Detection of Ultrasound 784 Part C Noncontact Methods Typically laser beams are used as the optical source, and these can provide monochromatic, linearly polarized, plane light beams. The electric field of such beams can be expressed as E = a exp[i(ωopt t − φ)] , (27.49) Part C 27.3 where E is the electric field of amplitude a, frequency ωopt , and phase φ. It is important to note that extant photodetectors cannot directly track the optical phase (the optical frequency ωopt is just too high), and as such only the optical intensity (proportional to P = E E ∗ = a2 , where ∗ represents the complex conjugate) can be directly measured. There are a number of ways in which ultrasound can affect the light beam. These can be broadly classified into intensity-modulated techniques and phase/frequency-modulated techniques. Intensity Modulation Induced by Ultrasound The intensity of the reflected light beam can change due to ultrasound-induced changes in the refractive index of the medium, and this can be monitored directly using a photodetector. Though these changes are typically very small for most materials, this method has been used successfully in picosecond ultrasonics [27.30] to measure the properties of thin films [27.31] and nanostructures [27.33]. Another intensity-based technique utilizes the surface tilt associated with ultrasonic motion [27.34]. The probe light beam is tightly focused onto an optically reflective object surface. A partial aperture (usually called a knife edge in this context) is placed behind a recollimating lens located in the path of the reflected beam prior to being focused onto a photodetector. The reflected light beam will undergo a slight tilt due to the ultrasonic displacement. This in turn will cause varying portions of light to be blocked by the knife edge, resulting in an intensity change at the photodetector. A third class of intensity-based techniques is applicable to continuous or tone-burst surface acoustic wave (SAW) packets of known frequency and velocity. In this case, the ultrasonic surface displacement acts like a surface diffraction grating, and an incident plane light beam will undergo diffraction in the presence of the SAW wave packet. A photodetector placed in the direction of either of the two expected diffracted first-order beams can be used to monitor the SAW wave packet [27.35]. Recently, diffraction detection has been used to measure the mechanical properties of thin films [27.36]. In general, intensitymodulated techniques are typically less sensitive than phase/frequency-modulated techniques. As such, their use in nondestructive characterization has been limited. The reader is referred to the review papers [27.4, 5] for further information on intensity-modulation techniques. Phase or Frequency Modulation Induced by Ultrasound Ultrasonic motion on the surface of a body also affects the phase or frequency of the reflected or scattered light. For simplicity, consider an object surface illuminated at normal incidence by a light beam, as shown in Fig. 27.16. Let the surface normal displacement at the point of measurement be u(t) due to ultrasonic motion, where t is time. We shall assume that the surface tilt is not so large that the reflected optical beam is tilted significantly away. Therefore the object surface displacement just changes the phase of the light by causing a change in the path length (equal to twice the ultrasonic normal displacement) that the light has to travel. In the presence of ultrasonic displacement, the electric field can therefore be expressed as E s = as exp[i(ωopt t − 2kopt u(t) − ϕs )] , (27.50) where kopt = 2π/λopt is the optical wavenumber, λopt is the optical wavelength, and ϕs is the optical phase (from some common reference point) in the absence of ultrasound. For time-varying phase modulation such as that caused by an ultrasonic wave packet, it is also possible to view the optical interaction with the surface motion as an instantaneous Doppler shift in optical frequency. To see this, note that (27.50) can be equivalently written in terms of the surface velocity V (t) = du/ dt as follows: E s = as exp[i(ω̃opt t − ϕs )] , (27.51) u (x1, x2, t) Incident light Reflected light Fig. 27.16 Phase modulation of light due to ultrasonic dis- placement Photoacoustic Characterization of Materials where the instantaneous optical frequency is now given by 't ω̃opt t = 2V ωopt 1 − dt , c (27.52) 0 27.3.2 Optical Interferometry The phase of a single optical beam cannot be measured directly since the optical frequency is too high to be monitored directly by any extant photodetector. Therefore, a demodulation scheme has to be used to retrieve phase-encoded information. There are a number of optical interferometers that perform this demodulation (see [27.37] for a general discussion on optical interferometry). Here we will only consider two systems that have found wide extensive application in photoacoustic metrology. Another common device that has found extensive applications in photoacoustic metrology is the confocal Fabry–Pérot interferometer. Due to space limitations, this will not be discussed here, and the reader is referred to the review paper by Dewhurst and Shan [27.6]. Reference-Beam Interferometers The simplest optical interferometer is the two-beam Michelson setup shown in Fig. 27.17. The output from a laser is split into two at a beam splitter and one of the beams is sent to the test object, and the other is sent to a reference mirror. Upon reflection, the two beams are recombined parallel to each other and made to interfere at the photodetector. The electric fields at the photodetector plane can be written as E r = ar exp[i(ωopt t − kopt L r )] , E s = as exp[i(ωopt t − kopt (L s + 2u(t)))] , (27.53) (27.54) where (i = r, s) refer to the reference and signal beams, respectively. Here E i are the electric fields of the two beams of amplitudes ai and optical frequency ωopt . The phases φi = kopt L i are due to the different path lengths L i that the two beams travel from a point of common phase (say at the point just prior to the two beams splitting at the beam splitter). Here, it is convenient to consider the phase term for the signal beam as being comprised of a static part kopt L s due to the static path length, and a time-varying part due to the time-varying ultrasonic displacement u(t). The total electric field at the photodetector plane is then the sum of the fields of the two beams, and the resulting intensity is therefore obtained as PD = Ptot {1 + M cos[kopt (L r − L s ) − 2kopt u(t)]} , (27.55) Pi = ai2 where are the optical intensities (directly proportional to the power in Watts) of the two beams individually. In the last expression above, we have defined the total optical power Ptot = Pr + Ps . The factor √ 2 Pr Ps M= Ptot is known as the modulation depth of the interference and ranges between 0 (when one of the beams is not present) to 1 (when the two beams are of equal intensity). If the phase change due to the signal of interest 2kopt u(t) 1 – as is the case for typical ultrasonic displacements – the best sensitivity is obtained by ensuring I Photodetector Laser Signal beam Reference beam Fig. 27.17 Two-beam homodyne (Michelson) interferome- ter 785 Δφ Fig. 27.18 Two-beam interferometer output intensity as a function of phase change. The largest variation in output intensity for small phase changes occurs at quadrature Part C 27.3 where c is the speed of light. The surface velocity associated with the ultrasonic motion therefore leads to a frequency shift of the optical beam. 27.3 Optical Detection of Ultrasound 786 Part C Noncontact Methods Part C 27.3 that the static phase difference is maintained at quadrature, i. e., at kopt (L r − L s ) = π2 (Fig. 27.18). This can be achieved by choosing the reference and signal beam path lengths appropriately. The two-beam Michelson interferometer that is maintained at quadrature therefore provides an output optical power at the photodetector given by PD = Ptot [1 + M(2kopt u(t))] , kopt u 1 . (27.56) This shows that the output of a Michelson interferometer that is operated at quadrature is proportional to the ultrasonic displacement. In reality, even the static optical path is not quite static because of low-frequency ambient vibration that can move the various optical components or the object around. If the signal of interest is high frequency (say several kHz or higher) – which is the case for ultrasonic signals – it is possible to use an active stabilization system using a moving mirror (typically mounted on a piezoelectric stack) on the reference leg such that the static (or, more appropriately, low-frequency) phase difference is always actively kept constant by means of a feedback controller. The piezoelectric mirror can also be used to calibrate the full fringe interferometric output by intentionally inducing an optical phase change in the reference leg that is larger than 2π. This will provide both the total power Ptot and the modulation depth M. Therefore, an absolute measurement of the ultrasonic displacement can be obtained from (27.56). It is important to characterize the signal-to-noise ratio (SNR), or equivalently the minimum detectable displacement of the optical interferometer. There are several possible noise sources in an optical detection system. These include noise from the laser source, in the photodetector, in the electronics, and noise in the a) Speckled object Photodetector Planar reference b) Speckled object Photodetector Wavefront-matched reference Fig. 27.19a,b Ultrasound detection on rough surfaces. (a) Interference of speckled signal and planar reference beams is nonoptimal. (b) Wavefront-matched interference optical path due to ambient vibrations and thermal currents. Most of these noise sources can be stabilized against or minimized by careful design and isolation, leaving only quantum or shot noise arising from random fluctuations in the photocurrent. Shot noise increases with increasing optical power, and it therefore basically sets the absolute limit of detection for optical measurement systems. It can be shown that the SNR of a shot-noise-limited Michelson interferometer operating at quadrature is given by [27.5] ( ηPtot (27.57) , SNR = kopt MU hνopt B where η is the detector quantum efficiency, νopt is the optical (circular) frequency in Hertz, h is Planck’s constant, B is the detection bandwidth, and U is the ultrasonic displacement. The minimum detectable ultrasonic signal can then be readily determined from (27.57) based on the somewhat arbitrary criterion that a signal is detectable if it is equal to the noise magnitude, i. e., if SNR = 1. For a detection bandwidth of 1 Hz, detector efficiency of 0.5, modulation depth M of 0.8, and total collected optical power of 1 W from a green laser (514 nm), the minimum detectable sensitivity is on the order of 10−17 m. Of all possible configurations, the two-beam homodyne interferometer provides the best shot-noise detection sensitivity as long as the object beam is specularly reflective. If the object surface is rough, the scattered object beam will in general be a speckled beam. In this case, the performance of the Michelson interferometer will be several orders of magnitude poorer due to two factors. First, the total optical power Ptot collected will be lower than from a mirror surface. Secondly, the mixing of a nonplanar object beam (one where the optical phase varies randomly across the beam) with a planar reference beam is not efficient (Fig. 27.19a), and indeed could be counterproductive with the worst-case situation leading to complete signal cancellation occurring. Therefore, interferometers such as the Michelson, which use a planar reference beam, are best used in the laboratory on optically mirror-like surfaces. For optically scattering surfaces, self-referential interferometers such as the adaptive interferometers based on two-wave mixing in photorefractive crystals are preferred. Dynamic Holographic Interferometers This class of interferometers is based on dynamic holographic recording typically in photorefractive media. Photoacoustic Characterization of Materials (a) creation of optical intensity gratings due to coherent interference of the interacting beams, leading to (b) nonuniform photoexcitation of electric charges in the PRC, which then diffuse/drift to create (c) a space-charge field within the PRC, which in turn creates (d) a refractive index grating via the electro-optical effect, and which causes (e) diffraction of the interacting beams A net consequence of this is that at the output of the PRC we have not only a portion of the transmitted probe beam, but also a part of the pump beam which is diffracted into the direction of the probe beam. The pump beam diffracted into the signal beam direction has the same wavefront structure as the transmitted signal beam. Since the PRC process has a certain response time (depending on the material, the applied electric field, and the total incident optical intensity), it turns out that it is unable to adapt to sufficiently high-frequency modulations in the signal beam. The PRC can only adapt to changes in the incident beams that are slower than the response time. This makes two-wave mixing in- λ/2 PRC BC λ/2 terferometers especially useful for ultrasound detection. High-frequency ultrasound-induced phase modulations are essentially not seen by the PRC, and therefore the diffracted pump beam will have the same wavefront structure as the unmodulated signal beam. The transmitted signal beam, however, obviously will contain the ultrasound-induced phase modulation. By interfering the diffracted (but unmodulated) pump beam with the transmitted (modulated) signal beam (both otherwise with the same wavefront structure) one obtains a highly efficient interferometer. Furthermore, any lowfrequency modulation in the interfering beams (such as those caused by noise from ambient vibration, or slow motion of the object) will be compensated for by the PRC as it adapts and creates a new hologram. Two-wave mixing interferometers therefore do not need any additional active stabilization against ambient noise. Several different types of photorefractive two-wave mixing interferometers have been described in the literature [27.40–42]. Here we will describe the isotropic diffraction configuration [27.44] shown in Fig. 27.20. For simplicity, optical activity and birefringence effects in the PRC will be neglected. Let the signal beam obtained from the scatter from the test object be s-polarized. As shown in Fig. 27.20, a half-wave plate (HWP) is used to rotate the incident signal beam polarization by 45◦ leading to both s- and p-polarized phase-modulated components of equal intensity given by a (27.58) E s0 = √ exp{i[ωopt t − φ(t)]} . 2 A photorefractive grating is created by the interference of the s-polarized component of the signal beam with the s-polarized pump beam. The diffracted pump beam (also s-polarized for this configuration of the PRC) upon exiting the crystal is then given by [27.44] a E r = √ exp(iωopt t) exp(−αL/2) 2 × {[exp(γ L) − 1] + exp[−iφ(t)]} , (27.59) BS Signal beam PD 2 Pump beam PD 1 Fig. 27.20 Configuration of isotropic diffraction setup (λ/2: half-wave plate, PRC: photorefractive crystal, BC: Berek compensator, BS: beam splitter; PD: photodetector) 787 Part C 27.3 One approach is to planarize the speckled object beam by using optical phase conjugation [27.38, 39]. The planarized object beam can then be effectively interfered with a planar reference beam in a two-beam homodyne or heterodyne interferometer. Alternatively, a reference beam with the same speckle structure as the static object beam can be holographically reconstructed for interference with the object beam containing the ultrasonic information (Fig. 27.19b) [27.40–42]. This is most readily achieved by using the process of two-wave mixing in photorefractive media [27.43]. Two-wave mixing is essentially a dynamic holographic process in which two coherent optical beams (pump/reference and probe/signal beams) interact within a photorefractive crystal (PRC). The process of TWM can be briefly summarized as 27.3 Optical Detection of Ultrasound 788 Part C Noncontact Methods a) Detection beam splitter (PBS) oriented at 45◦ to the s- and pdirections, giving rise to two sets of optical beams that interfere at the two photodetectors. The intensities recorded at the two photodetectors are then given by (for φ(t) π/2) Ptot −αL 2γr L + 2 e2γr L {cos(γi L − φL ) e e PD1 = 4 +φ(t) sin(γi L − φL ) + sin(γi L)} + 2φ(t) sin φL + 1 Ptot −αL 2γr L PD2 = − 2 e2γr L {cos(γi L − φL ) e e 4 +φ(t) sin(γi L − φL ) − sin(γi L)} − 2φ(t) sin φL + 1 , (27.61) Source 25 m 3 1 Part C 27.3 R R 30 mm 70 mm b) 0.32 R 0.24 0.16 S1 0.08 P1 P3 0 –0.08 –0.16 RR –0.24 10 20 30 40 50 60 Time (μm) Fig. 27.21a,b Photoacoustically generated waves detected on an un- polished aluminum block using a two-wave mixing interferometer. (a) Specimen configuration. (b) Optically detected ultrasonic signal. The vertical axis is displacement in nanometers (R: direct Rayleigh wave; RR: once reflected Rayleigh wave; P1: once reflected longitudinal wave; P3: three-times reflected longitudinal wave; S1: once reflected shear wave) where γ = γr + iγi is the complex photorefractive gain, α is the intensity absorption coefficient of the crystal, and L is the length of the crystal. The p-polarized component of the signal beam is transmitted by the PRC undisturbed (except for absorption) and may be written as a E s = √ exp(−αL/2) exp{i[ωopt t − φ(t)]} . 2 (27.60) Upon exiting the PRC, the diffracted pump and the transmitted signal beams are now orthogonally polarized. A Berek’s wave plate is interposed so as to introduce an additional phase shift of φL in the transmitted signal beam to put the interference at quadrature. The two beams are then passed through a polarizing where Ptot = a2 is proportional to the optical power collected in the scattered object beam. In the case of a pure real photorefractive gain, quadrature is obtained by setting φL = π/2. In this case, upon electronically subtracting the two photodetector signals using a differential amplifier, the output signal is (27.62) S = Ptot e−αL eγ L − 1 ϕ(t) . Since the phase modulation φ(t) = 2kopt u(t) , the output signal is directly proportional to the ultrasonic displacement. The signal-to-noise ratio of the two-wave mixing interferometer in the isotropic configuration for real photorefractive gain can be shown to be [27.44] ( ηPtot − αL eγ L − 1 e 2 SNR = 2kopt U 1/2 . hvopt e2γ L + 1 (27.63) It is clear that, the lower the absorption and the higher the photorefractive gain, the better the SNR. Minimum detectable sensitivities on the order of 10−15 m have been reported on optically scattering surfaces using dynamic holographic interferometers [27.42]. As an illustration, Fig. 27.21 shows the ultrasonic surface displacements monitored using a two-wave mixing interferometer on an unpolished aluminum block using a 500 mW laser source. The generation was using a photoacoustic point source using a 15 mJ Nd:YAG pulsed laser. Photoacoustic Characterization of Materials 27.3.3 Practical Considerations: Systems for Optical Detection of Ultrasound put together by an experienced engineer. Of greater interest are confocal Fabry–Pérot systems and the dynamic holographic interferometers, which work well on unpolished surfaces not only in the laboratory but also in industrial settings. These detection systems are commercially available from a number of sources. 27.4 Applications of Photoacoustics Photoacoustic methods have found wide-ranging applications in both industry and academic research. Here we will consider some illustrative applications in nondestructive flaw identification and materials characterization. 27.4.1 Photoacoustic Methods for Nondestructive Imaging of Structures Photoacoustic techniques have been used for nondestructive flaw detection in metallic and composite structures. Here we review a few representative example applications in flaw imaging using bulk waves, surface acoustic waves, and Lamb waves. Flaw Imaging Using Bulk Waves As discussed earlier, thermoelastic generation of bulk waves in the epicentral direction is generally weak in materials for which the optical penetration depth and thermal diffusion effects are small. Therefore, laser ultrasonic techniques using bulk waves have been used primarily for imaging of defects in composite structures (where the penetration depth is large), or on structures that are coated with a sacrificial film that enhances epicentral generation. Recently, Zhou et al. [27.45] have developed efficient photoacoustic generation layers which they have used in conjunction with an ultrasonic imaging camera to image the interior of aluminum and composite structures (Fig. 27.22). Lockheed Martin has recently installed a large-scale laser ultrasonic facility for inspecting polymer-matrix composite structures in aircraft such as the joint strike fighter [27.46]. In this system, a pulsed CO2 laser was used to thermoelastically generate bulk waves into the composite part. A coaxial long-pulse Nd:YAG detection laser demodulated by a confocal Fabry–Pérot was used to monitor the back reflections of the bulk waves. The system has been demonstrated on prototype F-22 inlet ducts. Yawn et al. [27.46] estimate that the inspection time using the noncontact laser system is about 70 min as opposed to about 24 h for a conventional ultrasonic squirter system. Flaw Imaging Using Surface Acoustic Waves Photoacoustic systems can also be used to generate and detect surface acoustic waves on specimens such as the aircraft wheel shown in Fig. 27.23a,b. Here the high de- Fig. 27.22 Imaging of subsurface features in aluminum structures using photoacoustically generated bulk waves. The subsurface features were 12.7 mm inside an aluminum block, and the feature size is on the order of 6.35 mm 789 Part C 27.4 Reference-beam interferometers, discussed in Sect. 27.3.2, are primarily laboratory tools which work well typically only on highly polished reflective specimens. These systems can be very easily 27.4 Applications of Photoacoustics 790 Part C Noncontact Methods a) b) EDM notch 15 mm 11 mm Part C 27.4 Laser generation Fiber-optic interferometer probe c) Surface displacement (nm) d) Surface displacement (nm) 0.15 0.15 Direct rayleigh 0.1 0.11 0.07 Reflection from crack 0.05 0.03 0 –0.05 –0.01 0 2 4 6 8 –0.05 10 Time (μs) 0 2 4 6 8 10 Time (μs) e) Reflection coefficient 0.5 0.4 0.3 0.2 0.1 0 –0.1 –5 Actual crack position 0 5 10 15 20 25 Position (mm) Fig. 27.23a–e Photoacoustic surface acoustic wave imaging. (a) Aircraft wheel part containing cracks. (b) Schematic of the scanning setup. Signal from locations (c) without a crack and (d) with a crack indicated by presence of reflections. (e) Reflection coefficient measured at different scanning locations of the wheel gree of double curvature of the wheel makes the use of contact transducers difficult. Huang et al. [27.47] used laser generation along with dual-probe heterodyne interferometer detection. The presence of cracks along Photoacoustic Characterization of Materials Tomographic Imaging Using Lamb Waves Tomographic imaging of plate structures using Lamb waves is often desired when the test area is not directly accessible and so must be probed from outside the area. Computer algorithms are used to reconstruct variations of a physical quantity (such as ultrasound attenuation) within a cross-sectional area from its integrated projection in all directions across that area. Photoacoustic tomographic systems using attenuation of ultrasound for tomographic reconstruction need to take into account the high degree of variability in the generated ultrasound arising from variation in the thermal absorption at different locations on the plate. A schematic of the setup is shown in Fig. 27.24a. Narrow-band Lamb waves were generated using an array of ther- moelastic sources. Figure 27.24b shows a cross-section of the simulated corrosion defect produced in epoxybonded aluminum plates. The specimen is composed of two aluminum plates of thickness 0.65 mm and an approximately 13 μm-thick epoxy film. Corrosion was simulated by partially removing the surface of the bottom plate and inserting a fine nickel powder in the cavity prior to bonding. Figure 27.24c,d show typical narrow-band Lamb waves detected using the dual-probe interferometer in the presence and absence of the inc) Amplitude (mV) 0.2 0.15 0.1 0.05 0 –0.05 –0.1 –0.15 –0.2 0 5 10 15 Time (μs) 5 10 15 Time (μs) d) Amplitude (mV) 0.2 0.15 a) 0.1 YAG Laser 0.05 Photodiode 0 Dual-probe fiber-optic interferometer –0.05 –0.1 Defects Specimen Oscilloscope b) Computer display –0.15 GPIB Line source Rotation and translation stages 5.5 mm 0.65 mm 0.65 mm Aluminum plates 0.22 mm Ni powder Bonded together by epoxy film Fig. 27.24a–e Photoacoustic Lamb-wave tomography. (a) Setup. (b) Cross-section of bonded thin plates con- taining an inclusion. Typical Lamb-wave signals after band-pass filtering (c) without defect, (d) with the defect between the two detecting points. (e) Superposed image of the tomographic image (solid line) and a conventional ultrasonic C-scan image –0.2 e) 0 791 Part C 27.4 the doubly curved location is indicated by the presence of reflected ultrasound signals in some locations but not in others (Fig. 27.23c,d). A reflection coefficient, calculated as the ratio of the reflected signal to the original signal power, is plotted in Fig. 27.23e and provides a measure of the length of the crack. Such pitch-catch approaches to detecting cracks using laser ultrasonics are feasible on cracks that are large enough to provide a sufficiently strong reflected signal. 27.4 Applications of Photoacoustics 792 Part C Noncontact Methods Part C 27.4 clusion. By tomographically scanning the plate, Nagata et al. [27.48] were able to create a tomographic image of the specimen, as shown in Fig. 27.24e. Also shown superposed in Fig. 27.24e is an ultrasonic C-scan image of the same sample obtained using a commercial scanning acoustic microscope. The size and the shape of the tomographically reconstructed image is consistent with that of the C-scan. Scanning Laser Source Imaging of Surface-Breaking Flaws In the applications described thus far, photoacoustic methods have been used in a conventional pitch-catch ultrasonic inspection mode, except that lasers were used to generate and detect the ultrasound. For detecting very small cracks, the pitch-catch technique requires that the crack reflect a significant fraction of the in- Laser Receiver source a) x2 Scanning Crack x1 x3 b) Interferometer signal (mV) 4 4 4 3 3 3 2 2 2 1 1 1 0 0 0 –1 –1 –1 –2 0 0.5 1 1.5 2 –2 2.5 3 Time (μs) 0.5 0 1 1.5 2 2.5 3 Time (μs) –2 0 0.5 1 1.5 2 2.5 3 Time (μs) c) Ultrasonic amplitude (mV) 6 II 5 Crack 4 I 3 III 2 1 0 0 1 2 3 4 5 6 SLS position (mm) Fig. 27.25a–c The scanning laser source (SLS) technique. (a) Schematic of the SLS technique. (b) Ultrasonic surface normal displacement as a function of time (in μs) at three locations of the scanning laser source – left: far from the defect; center: close to the defect; and right: behind the defect. (c) Typical characteristic signature of ultrasonic peak-to-peak amplitude versus SLS location as the source is scanned over a surface-breaking defect Photoacoustic Characterization of Materials 793 tering of the generated signal by the defect (zone III in Fig. 27.25c). The variation in signal amplitude is due to two mechanisms: (a) near-field scattering by the defect (b) changes in the conditions of generation of ultrasound when the SLS is in the vicinity of the defect As such, the SLS technique is not very sensitive to flaw orientation [27.50]. In addition to the amplitude signature shown above, spectral variations in the detected ultrasonic signal also show characteristic features [27.50]. Both amplitude and spectral variations can form the basis for an inspection procedure using the SLS technique. The SLS technique has been applied to detect small electric discharge machined notches in titanium engine disk blade attachment slots [27.49]. 27.4.2 Photoacoustic Methods for Materials Characterization Photoacoustic metrology has been used to characterize the properties of materials ranging from the macroto the microscale. Here we describe applications to thin-film and coating characterization and to the determination of material anisotropy. Characterization of Material Anisotropy Photoacoustic methods have been used by a number of researchers to investigate the anisotropy of materials [27.51–55]. Both point- and line-focused laser Specimen y Circular array receiver r θ Point source x Fig. 27.26 Point source generation and multiplexed array detection of surface acoustic waves in anisotropic materials Part C 27.4 cident wave, and furthermore that the generating and receiving locations be in line and normal to the crack. Recently, Kromine et al. [27.49, 50] have developed a scanning laser source (SLS) technique for detecting very small surface-breaking cracks that are arbitrarily oriented with respect to the generating and detecting directions. The SLS technique has no counterpart in conventional ultrasonic inspection methodologies as it relies on near-field scattering and variations in thermoelastic generation of ultrasound in the presence and absence of defects. In the SLS technique, the ultrasound generation source, which is a point- or line-focused high-power laser beam, is swept across the test specimen surface and passes over surface-breaking flaws (Fig. 27.25a). The generated ultrasonic wave packet is detected using an optical interferometer or a conventional contact piezoelectric transducer either at a fixed location on the specimen or at a fixed distance between the source and receiver. The ultrasonic signal that arrives at the Rayleigh wave speed is monitored as the SLS is scanned. Kromine et al. [27.50] and Sohn et al. [27.15] have shown that the amplitude and frequency of the measured ultrasonic signal have specific variations when the laser source approaches, passes over, and moves behind the defect. Kromine et al. [27.50] have experimentally verified the SLS technique on an aluminum specimen with a surface-breaking fatigue crack of 4 mm length and 50 μm width. A broadband heterodyne interferometer with 1–15 MHz bandwidth was used as the ultrasonic detector. The SLS was formed by focusing a pulsed Nd:YAG laser beam (pulse duration 10 ns, energy 3 mJ). The detected ultrasonic signal at three locations of the SLS position is presented in Fig. 27.25b. The Rayleigh wave amplitude as a function of the SLS position is shown in Fig. 27.25c. Several revealing aspects of the Rayleigh wave amplitude signature should be noted. In the absence of a defect or when the source is far ahead of the defect, the amplitude of the generated ultrasonic direct signal is constant (zone I in Fig. 27.25c). The Rayleigh wave signal is of sufficient amplitude above the noise floor to be easily measured by the laser detector (Fig. 27.25b), but the reflection is within the noise floor. As the source approaches the defect, the amplitude of the detected signal significantly increases (zone II in Fig. 27.25c). This increase is readily detectable even with a low sensitivity laser interferometer as compared to weak echoes from the flaw (Fig. 27.25b). As the source moves behind the defect, the amplitude drops lower than in zone I due to scat- 27.4 Applications of Photoacoustics 794 Part C Noncontact Methods Part C 27.4 sources have been used to generate the ultrasound. The ultrasound generated by a point laser source is typically detected by a point receiver and group velocity information is obtained at different angles. Doyle and Scala [27.56] have used a line-focused laser to determine the elastic constants of composite materials using the measured phase velocities of surface acoustic waves. Huang and Achenbach [27.57] used a line source and a dual-probe Michelson interferometer to provide accurate measurements of time of flight of SAWs on silicon. Zhou et al. [27.58] have recently used a multiplexed two-wave mixing interferometer with eight detection channels to provide group velocity slowness images. Figure 27.26 shows the configuration of the optical beams for anisotropic material characterization using a) θ = 90° θ = 80° θ = 70° θ = 60° θ = 50° θ = 40° θ = 30° θ = 20° θ = 10° θ = 0° 0.8 1 1.2 b) × 10–4 1.4 1.6 1.8 Time (μs) 120 Slowness (ms/mm) 0.3 90 2.1 SAWs. In this setup, a pulsed Nd:YAG laser (pulse energy of approximately 1 mJ) is focused by a lens system onto the sample surface to generate the SAWs. The eight optical probe beams are obtained using a circular diffraction grating and focused onto the sample surface by a lens system to fall on a circle of radius r centered about the generation spot. The whole array of eight points was rotated every 2◦ to obtain the material anisotropy over the entire 360◦ range. Since a point source and point receiver configuration is used, the surface wave group velocity is obtained in this case. Figure 27.27a shows the time-domain SAW signals on (001) silicon from 0◦ to 90◦ . The group velocity slowness in each direction is obtained from the timedomain data through a cross-correlation technique and is shown in Fig. 27.27b where the filled circles are the experimental values. Also shown in Fig. 27.27b is the theoretical group velocity slowness calculated using nominal material values. The discontinuities that appear in both the experimental and the theoretical curves are due to the presence of pseudo-surface waves. Zhou et al. [27.58] have also obtained group velocity slowness curves on the (0001) surface of a block of quartz. The time-domain traces and the corresponding group velocity slowness curves are shown in Fig. 27.28. The multiple pulses observed are due to a combination of the presence of SAWs and pseudo-SAWs as well as the energy folding that occurs in anisotropic materials such as (0001) quartz. The group velocity slowness curves obtained experimentally can be further processed to ob- 60 Experiment Theory 0.2 2 150 30 0.1 1.9 0 1.8 180 0 –0.1 1.9 210 330 –0.2 2 –0.3 2.1 240 300 270 Fig. 27.27 (a) SAW signals detected in different directions on z-cut silicon. (b) Slowness curve for z-cut silicon –0.3 –0.2 –0.1 0 0.1 0.2 0.3 Slowness (ms/mm) Fig. 27.28 Group velocity slowness of z-cut quartz Photoacoustic Characterization of Materials tain the anisotropic material constants as described by Castagnede et al. [27.59]. PZT driver Attenuator 780 nm femtosecond laser λ/2 532 nm CW laser Aperture PBS λ/4 Stabilizer PZT BS Sample Optical fibers Oscilloscope BPD Fig. 27.29 Guided-wave photoacoustic setup (BS: beam splitter, PBS: polarized beam splitter, BPD: balanced photodetector, λ/4: quarter-wave plate, λ/2: half-wave plate, PZT: piezoelectric mounted mirror) of hundreds of nanometers, followed by a Cr adhesion metallic layer with thickness about 100 nm right on the steel substrate. The properties of the interpolated layer are taken to be the average of those of the DLC and the metallic layers. The transfer matrix method described in Sect. 27.1.3, was used to obtain the theoretical guided-wave dispersion curves. To derive the mechanical properties, an inverse problem has to be solved to calculate the parameters from the measured velocity dispersion curve v( f ). A nonlinear regression method is used to minimize the least-square error function y= N 2 1 theo v − vmeas , N 795 (27.64) i=1 where vmeas are the measured velocities, and vtheo are the theoretically calculated velocities, which are functions of the mechanical properties of each layer. A simplex method of least-square curve fitting is useful for fitting a function of more than one variable [27.64]. The thicknesses of the coatings were separately measured and used in the calculation. The reliability of the results mainly depends on the accuracy of the experiments, the choice of initial parameters, and the number of fitted variables. The Young’s modulus, Poisson’s ratio, and density of the DLC coating layer were set as variables to be determined in the iteration. Repetitive fitting showed a variation of up to 5% for the fitted values of Young’s moduli and densities. Part C 27.4 Characterization of the Mechanical Properties of Coatings The small footprint and noncontact nature of photoacoustic methods make them especially useful for characterizing coatings. Several optical techniques have been devised and implemented. A pump-probe technique has been used in which very high-frequency (GHz) acoustic waves are generated that propagate perpendicular to the film and reflect off of the film/substrate interface [27.26, 31]. This bulk wave technique requires an ultrafast laser source, and material attenuation of high-frequency ultrasound limits the useful measurement range to reasonably thin films. For thicker films, guided-wave ultrasonic techniques are more practical. The impulsive stimulated thermal scattering (ISTS) [27.60] technique and the phase velocity scanning (PVS) [27.61] technique both use a spatially periodic irradiance pattern to generate single-frequency surface acoustic wave (SAW) tone bursts which are detected through probe-beam diffraction, interferometry, or contact transducers. Broadband techniques [27.19, 62] can also be used where SAWs are generated with a simple pulsed laser point or line source which are then detected with an interferometer after some propagation distance along the film. Ultrahard coatings such as diamond, diamond-like carbon (DLC), cubic BN, etc., are of great interest due to their unique mechanical, thermal, and electrical properties. In particular, the high hardness and stiffness of diamond-like thin films make them excellent coating materials for tribological applications. Unfortunately, the coating properties are highly sensitive to the processing parameters, and photoacoustic methods provide a way to measure the properties of these coatings nondestructively. Figure 27.29 shows the guided-wave photoacoustic setup used for characterizing multilayer Cr-DLC specimens. A pulsed laser was line-focused to a line width of 10 μm on the surface of the specimen to generate broadband guided acoustic waves. The acoustic waves were monitored by a stabilized balanced Michelson interferometer. By monitoring the guided waves at multiple source to receiver locations (Fig. 27.30a), the dispersion curves for these waves were obtained, as shown in Fig. 27.30b. To interpret the measurements, the multilayer DLC specimen was modeled as a three-layer system [27.63]. Below the top layer of DLC, there is an interpolated transition layer of Cr and DLC with various thickness 27.4 Applications of Photoacoustics 796 Part C Noncontact Methods Part C 27.4 The experimentally measured dispersion curves for several Cr-DLC coatings are shown in Fig. 27.30b as dots. The solid lines in Fig. 27.30b are the theoretically fitted dispersion curves as described above for the inverse problem. The good agreement indicates that material properties can be extracted with a high degree of confidence. Note that significant variations in DLC a) Signal amplitude (arb. units) 0.2 0.1 0 –0.1 –0.2 –0.3 3 3.2 3.4 3.6 3.8 0.1 0 –0.1 –0.2 5 5.2 5.4 5.6 5.8 6 Time (μs) b) Phase velocity (m/s) 3060 (A) Cr-DLC specimens #1 #2 #3 #4 #5 #6 3040 3020 properties can be obtained, arising from fabrication process variations [27.63]. Photoacoustic characterization of such coatings is therefore very useful in process control applications to assure quality. Characterization of the Mechanical Properties of Thin Films Photoacoustic techniques can also be used to characterize the properties of free-standing nanometer-sized thin films [27.36, 65]. Such thin films are widely used in micro-electromechanical systems (MEMS) devices such as radiofrequency (RF) switches, pressure sensors, and micromirrors. Described below are guided-wave and bulk-wave photoacoustic methods for characterizing free-standing thin films. Two-layer thin films of Al/Si3 N4 were fabricated on a standard Si wafer using standard microfabrication processes [27.65]. The film thicknesses were in the range of hundreds of nanometers. Several windows were etched in the wafer to provide unsupported membranes of Al/Si3 N4 which were only edge supported (Fig. 27.31). In a first set of experiments, bulk-wave photoacoustic measurements were made. A standard pump-probe optical setup operated at 780 nm using a Ti:sapphire femtosecond laser (100 fs pulse duration, 80 MHz repetition rate) was used in this work. Both the substrate-supported and unsupported region were measured as, indicated in Fig. 27.31. Figure 27.32 shows the normalized measured pump-probe signals for the thin films on the silicon substrate supported region. As shown in Fig. 27.32, the time of flight of the first ultrasonic echo reflected from the first Al/Si3 N4 interface is marked as τ1 and that of the echo from the Si3 N4 /Si interface is marked as τ2 . To deduce Pump pulses Probe pulses 3000 Al Si3N4 2980 GW 0 d1 d2 BW Si 2960 0 50 100 x 150 Frequency (MHz) Fig. 27.30a,b Guided-wave photoacoustic measurement of Cr-DLC coatings. (a) Broadband SAW signals detected at two source–receiver locations. (b) SAW dispersion curves Michelson interferometer Fig. 27.31 Bulk-wave and guided-wave photoacoustic characterization of free-standing thin films: schematic diagram of the compact optical setup for both bulk-wave and guided-wave detection Photoacoustic Characterization of Materials 0.2 2 Short source –receiver distance 0.1 τ1 τ2 0 5 4 1 3 0.5 2 1 0 50 100 150 200 250 300 350 Delay time (ps) Fig. 27.32 The experimental transient thermoreflectivity Part C 27.4 Sample 6 0 797 a) Relative amplitude (arb. units) Reflection coefficient change (arb. units) 1.5 27.4 Applications of Photoacoustics – 0.1 – 0.2 – 0.3 Long source – receiver distance – 0.4 0.2 0.8 1 1.2 1.4 Time (μs) 1.6 Specimen #1 Specimen #2 Specimen #3 Specimen #4 1.4 where h is the thickness, ρ is the density of the film, and σ is the in-plane stress in the film (typically caused by residual stress). The flexural rigidity D is related to the Young’s modulus E, Poisson’s ratio ν, and the geometry of the film. 0.6 b) (phase velocity υa0)2 (km2/s2) signals for double-layer thin films the elastic moduli E accurately, the theoretical simulated transient thermoelastic signals (Sect. 27.2.2) with various moduli were calculated and compared with the experimental signals, and the moduli that give the smallest error for the time of flight of acoustic echoes were determined iteratively. In general, the measured Young’s modulus of the aluminum layer falls in the range 47–65 GPa. The experimentally determined Young’s moduli of silicon nitride films range from about 220 to 280 GPa and are in good agreement with the Young’s modulus of low-pressure chemical vapor deposition (LPCVD)-fabricated silicon nitride reported in the literature. The same set of specimens was also tested using guided-wave photoacoustics. Only the lowest-order antisymmetric Lamb-wave mode was efficiently generated and detected in such ultrathin films. For ultrathin films such as these, for small wavenumbers, a simple expression for the acoustic phase velocity of A0 mode in terms of the wavenumber k can be derived [27.65]: ( D 2 σ ω (27.65) k + , v A0 = = k ρh ρ 0.4 1.2 1 0.8 0.6 0.4 0.2 0 0 0.1 0.2 0.3 0.4 (k h)2 Fig. 27.33a,b Guided-wave photoacoustic characterization of two-layer ultrathin films. (a) Measured signals at two source-to-receiver locations for thin-film sample #1, and (b) measured dispersion curves (dots) and linear fitted curves (lines) for samples #1–#4. The average measured flexural rigidities are 4.82, 5.82, 3.64, 1.76 × 10−9 Nm; and the residual stresses are 235, 299, 334, 242 MPa, respectively Figure 27.33a shows typical measured time traces of the A0 mode at two source-to-receiver positions. Figure 27.33b shows the experimentally determined A0 mode dispersion curves for four specimens. From the figure, it is clear that (27.65) represents the measurements well, thereby enabling direct determination of the residual stresses in the film [27.65]. 798 Part C Noncontact Methods 27.5 Closing Remarks Part C 27 In this chapter, only a selective review of photoacoustic metrology as applied to mechanical characterization of solids has been provided. Applications of photoacoustic methods of course extend well beyond the ones discussed here. Photoacoustic spectroscopy is a wellestablished set of methods for the characterization of the composition of condensed and gaseous matter [27.66]. Recent developments in biomedical photoacoustics promise new methods of characterization and imaging of tumors and blood vessels [27.67, 68]. Photoacoustic methods are also being used to characterize nanostructures such as superlattices [27.69] and nanoparticles [27.70]. The photoacoustic phenomenon, which Alexander Graham Bell originally investigated, possibly with applications to telephony in mind, has since yielded several powerful metrology tools for both the laboratory researcher and for industrial applications. References 27.1 27.2 27.3 27.4 27.5 27.6 27.7 27.8 27.9 27.10 27.11 27.12 27.13 27.14 27.15 C.B. Scruby, L.E. Drain: Laser Ultrasonics: Techniques and Application (Adam Hilder, New York 1990) V.E. Gusev, A.A. Karabutov: Laser Optoacoustics (American Institute of Physics, New York 1993) D.A. Hutchins: Ultrasonic Generation by Pulsed Lasers. In: Physical Acoustics, Vol. XVIII, ed. by W.P. Mason, R.N. Thurston (Academic, New York 1988) pp. 21–123 J.P. Monchalin: Optical detection of ultrasound, IEEE Trans. Ultrason. Ferroelectr. Frequ. Control, 33(5), 485–499 (1986) J.W. Wagner: Optical Detection of Ultrasound. In: Physical Acoustics, Vol. XIX, ed. by R.N. Thurston, A.D. Pierce (Academic, New York 1990) R.J. Dewhurst, Q. Shan: Optical remote measurement of ultrasound, Meas. Sci. Technol. 10, R139–R168 (1999) J.D. Achenbach: Wave Propagation in Elastic Solids (North-Holland/Elsevier, Amsterdam 1973) B.A. Auld: Acoustic Fields and Waves in Solids (Krieger, Malabar 1990) D. Royer, E. Dieulesaint: Elastic Waves in Solids (Springer, New York 1996) G.W. Farnell: Properties of Elastic Surface Waves. In: Physical Acoustics, Vol. VI, ed. by W.P. Mason, R.N. Thurston (Academic, New York 1970) G.W. Farnell, E.L. Adler: Elastic Wave Propagation in Thin Layers. In: Physical Acoustics, Vol. IX, ed. by W.P. Mason, R.N. Thurston (Academic, New York 1974) W.T. Thomson: Transmission of Elastic Waves through a Stratified Solid Medium, J. Appl. Phys. 21, 89–93 (1950) N.A. Haskell: The dispersion of surface waves on multilayered media, Bull. Seism. Soc. Am. 43(1), 17–34 (1953) R.M. White: Generation of elastic waves by transient surface heating, J. Appl. Phys. 24(12), 3559–3567 (1963) Y. Sohn, S. Krishnaswamy: Interaction of a scanning laser-generated ultrasonic line source with 27.16 27.17 27.18 27.19 27.20 27.21 27.22 27.23 27.24 27.25 27.26 27.27 a surface-breaking flaw, J. Acoust. Soc. Am. 115(1), 172–181 (2004) A.M. Aindow, D.A. Hutchins, R.J. Dewhurst, S.B. Palmer: Laser-generated ultrasonic pulses at free metal-surfaces, J. Acoust. Soc. Am. 70, 449– 455 (1981) R.J. Dewhurst, C. Edwards, A.D.W. McKie, S.B. Palmer: Estimation of the thickness of thin metal sheet using laser generated ultrasound, Appl. Phys. Lett. 51, 1066–1068 (1987) J. Spicer, A.D.W. McKie, J.W. Wagner: Quantitative theory for laser ultrasonic waves in a thin plate, Appl. Phys. Lett. 57(18), 1882–1884 (1990) T.W. Murray, S. Krishnaswamy, J.D. Achenbach: Laser generation of ultrasound in films and coatings, Appl. Phys. Lett. 74(23), 3561–3563 (1999) T.W. Murray, J.W. Wagner: Laser generation of acoustic waves in the ablative regime, J. Appl. Phys. 85(4), 2031–2040 (1999) F.A. McDonald: Practical quantitative theory of photoacoustic pulse generation, Appl. Phys. Lett. 54(16), 1504–1506 (1989) J. Spicer: Laser Ultrasonics in Finite Structures: Comprehensive Modeling with Supporting Experiments. Ph.D. Thesis (The Johns Hopkins University, Baltimore 1991) I. Arias: Modeling of the Detection of Surface Breaking Cracks by Laser Ultrasonics. Ph.D. Thesis (Northwestern University, Evanston 2003) C.B. Scruby, R.J. Dewhurst, D. Hutchins, S. Palmer: Quantitative studies of thermally-generated elastic wave in laser irradiated solids, J. Appl. Phys. 51, 6210–6216 (1980) L. Rose: Point-source representation for laser generated ultrasound, J. Acoust. Soc. Am. 75(3), 723–732 (1984) C. Thomsen, H.T. Grahn, H.J. Maris, J. Tauc: Surface generation detection of phonons by picosecond light-pulses, Phys. Rev. B 34, 4129–4138 (1986) C.J.K. Richardson, M.J. Ehrlich, J.W. Wagner: Interferometric detection of ultrafast thermo-elastic Photoacoustic Characterization of Materials 27.28 27.29 27.31 27.32 27.33 27.34 27.35 27.36 27.37 27.38 27.39 27.40 27.41 27.42 27.43 27.44 27.45 27.46 27.47 27.48 27.49 27.50 27.51 27.52 27.53 27.54 27.55 P. Delaye, A. Blouin, D. Drolet, L.A. Montmorillon, G. Roosen, J.P. Monchalin: Detection of ultrasonic motion of a scattering surface by photorefractive InP:Fe under an applied dc field, J. Opt. Soc. Am. B 14(7), 1723–1734 (1997) Y. Zhou, G. Petculescu, I.M. Komsky, S. Krishnaswamy: A high-resolution, real-time ultrasonic imaging system for NDI applications, SPIE, Vol. 6177 (2006) K.R. Yawn, T.E. Drake, M.A. Osterkamp, S.Y. Chuang, D. Kaiser, C. Marquardt, B. Filkins, P. Lorraine, K. Martin, J. Miller: Large-Scale Laser Ultrasonic Facility for Aerospace Applications. In: Review of Progress in Quantitative Nondestructive Evaluation, Vol. 18, ed. by D.O. Thompson, D.E. Chimenti (Kluwer Academic/Plenum, New York 1999) pp. 387–393 J. Huang, Y. Nagata, S. Krishnaswamy, J.D. Achenbach: Laser based ultrasonics for flaw detection, IEEE Ultrasonic Symposium, ed. by M. Levy, S.C. Schneider (IEEE, New York 1994) pp. 1205– 1209 Y. Nagata, J. Huang, J.D. Achenbach, S. Krishnaswamy: Computed Tomography Using LaserBased Ultrasonics. In: Review of Progress in Quantitative Nondestructive Evaluation, Vol. 14, ed. by D.O. Thompson, D.E. Chimenti (Plenum, New York 1995) A. Kromine, P. Fomitchov, S. Krishnaswamy, J.D. Achenbach: Scanning Laser Source Technique and its Applications to Turbine Disk Inspection. In: Review of Progress in Quantitative Nondestructive Evaluation, Vol. 18A, ed. by D.O. Thompson, D.E. Chimenti (Plenum, New York 1998) pp. 381– 386 A.K. Kromine, P.A. Fomitchov, S. Krishnaswamy, J.D. Achenbach: Laser ultrasonic detection of surface breaking discontinuities: Scanning laser source technique, Mater. Eval. 58(2), 173–177 (2000) A.A. Maznev, A. Akthakul, K.A. Nelson: Surface acoustic modes in thin films on anisotropic substrates, J. Appl. Phys. 86(5), 2818–2824 (1999) D.C. Hurley, V.K. Tewary, A.J. Richards: Surface acoustic wave methods to determine the anisotropic elastic properties of thin films, Meas. Sci. Technol. 12(9), 1486–1494 (2001) B. Audoin, C. Bescond, M. Deschamps: Measurement of stiffness coefficients of anistropic materials from pointlike generation and detection of acoustic waves, J. Appl. Phys. 80(7), 3760–3771 (1996) A. Neubrand, P. Hess: Laser generation and detection of surface acoustic waves: elastic properties of surface layers, J. Appl. Phys. 71(1), 227–238 (1992) A.G. Every, W. Sachse: Determination of the elastic constants of anisotropic solids from acoustic-wave group-velocity measurements, Phys. Rev. B 42, 8196–8205 (1990) 799 Part C 27 27.30 transients in thin films: theory with supporting experiment, J. Opt. Soc. Am. B 16, 1007–1015 (1999) Z. Bozoki, A. Miklos, D. Bicanic: Photothermoelastic transfer matrix, Appl. Phys. Lett. 64, 1362–1364 (1994) A. Miklos, Z. Bozoki, A. Lorincz: Picosecond transient reflectance of thin metal films, J. Appl. Phys. 66, 2968–2972 (1989) H.T. Grahn, H.J. Maris, J. Tauc: Picosecond ultrasonics, IEEE J. Quantum Electron. 25(12), 2562–2569 (1989) C. Thomsen, H.T. Grahn, H.J. Maris, J. Tauc: Ultrasonic experiments with picosecond time resolution, J. Phys. Colloques. C10 46, 765–772 (1985) P.A. Fleury: Light scattering as a probe of phonons and other excitations. In: Physical Acoustics, Vol. VI, ed. by W.P. Mason, R.N. Thurston (Academic, New York 1970) G.A. Antonelli, H.J. Maris, S.G. Malhotra, M.E. Harper: Picosecond ultrasonics study of the vibrational modes of a nanostructure, J. Appl. Phys. 91(5), 3261–3267 (2002) R. Adler, A. Korpel, P. Desmares: An instrument for making surface waves visible, IEEE Trans. Sonics. Ulrason. SU-15, 157–160 (1967) E.G.H. Lean, C.C. Tseng, C.G. Powell: Optical probing of acoustic surface-wave harmonic generation, Appl. Phys. Lett. 16(1), 32–35 (1970) J.A. Rogers, G.R. Bogart, R.E. Miller: Noncontact quantitative spatial mapping of stress and flexural rigidity in thin membranes using a picosecond transient grating photoacoustic technique, J. Acoust. Soc. Am. 109, 547–553 (2001) D. Malacara: Optical Shop Testing (Wiley, New York 1992) M. Paul, B. Betz, W. Arnold: Interferometric detection of ultrasound from rough surfaces using optical phase conjugation, Appl. Phys. Lett. 50, 1569–1571 (1987) P. Delaye, A. Blouin, D. Drolet, J.P. Monchalin: Heterodyne-detection of ultrasound from rough surfaces using a double-phase conjugate mirror, Appl. Phys. Lett. 67, 3251–3253 (1995) R.K. Ing, J.P. Monchalin: Broadband optical detection of ultrasound by two-wave mixing in a photorefractive crystal, Appl. Phys. Lett. 59, 3233–3235 (1991) P. Delaye, L.A. de Montmorillon, G. Roosen: Transmission of time modulated optical signals through an absorbing photorefractive crystal, Opt. Commun. 118, 154–164 (1995) B.F. Pouet, R.K. Ing, S. Krishnaswamy, D. Royer: Heterodyne interferometer with two-wave mixing in photorefractive crystals for ultrasound detection on rough srufaces, Appl. Phys. Lett. 69, 3782–2784 (1996) P. Yeh: Introduction to Photorefractive Nonlinear Optics (Wiley, New York 1993) References 800 Part C Noncontact Methods 27.56 27.57 Part C 27 27.58 27.59 27.60 27.61 27.62 27.63 P.A. Doyle, C.M. Scala: Toward laser-ultrasonic characterization of an orthotropic isotropic interface, J. Acoust. Soc. Am. 93(3), 1385–1392 (1993) J. Huang, J.D. Achenbach: Measurement of material anisotropy by dual-probe laser interferometer, Res. Nondestr. Eval. 5, 225–235 (1994) Y. Zhou, T.W. Murray, S. Krishnaswamy: Photoacoustic imaging of surface wave slowness using multiplexed two-wave mixing interferometry, IEEE UFFC 49(8), 1118–1123 (2002) B. Castagnede, K.Y. Kim, W. Sachse, M.O. Thompson: Determination of the elastic constants of anistropic materials using laser-generated ultrasonic signals, J. Appl. Phys. 70(1), 150–157 (1991) A.R. Duggal, J.A. Rogers, K.A. Nelson: Real-time optical characterization of surface acoustic modes of polyimide thin-film coatings, J. Appl. Phys. 72(7), 2823–2839 (1992) A. Harata, H. Nishimura, T. Sawada: Laser-induced surface acoustic-waves and photothermal surface gratings generated by crossing two pulsed laserbeams, Appl. Phys. Lett. 57(2), 132–134 (1990) D. Schneider, T. Schwartz, A.S. Bradfordm, Q. Shan, R.J. Dewhurst: Controlling the quality of thin films by surface acoustic waves, Ultrasonics 35, 345–356 (1997) F. Zhang, S. Krishnaswamy, D. Fei, D.A. Rebinsky, B. Feng: Ultrasonic characterization of mechanical 27.64 27.65 27.66 27.67 27.68 27.69 27.70 properties of diamond-like carbon hard coatings, Thin Solid Films 503, 250–258 (2006) J.D. Achenbach, J.O. Kim, Y.C. Lee: Acoustic microscopy measurement of elastic-constants and mass density. In: Advances in Acoustic Microscopy, ed. by A. Briggs (Plenum, New York 1995) C.M. Hernandez, T.W. Murray, S. Krishnaswamy: Photoacoustic characterization of the mechanical properties of thin films, Appl. Phys. Lett. 80(4), 691–693 (2002) A. Rosencwaig: Photoacoustics and photoacoustic spectroscopy (Wiley, New York 1980) R.G.M. Kolkman, E. Hondebrink, W. Steenburgen, F.F.M. de Mul: In vivo photoacoustic imaging of blood vessels using an extreme-narrow aperture sensor, IEEE J. Sel. Top. Quantum Electron. 9(2), 343–346 (2003) V.G. Andreev, A.A. Karabutov, A.A. Oraevsky: Detection of ultrawide-band ultrasound pulses in optoacoustic tomography, IEEE Trans. Ultrason, Ferroelectr. Frequ. Control 50(10), 1183–1390 (2003) N.W. Pu: Ultrafast excitation and detection of acoustic phonon modes in superlattices, Phys. Rev. B 72, 115428 (2005) M. Nisoli, S. de Silvestri, A. Cavalleri, A.M. Malvezzi, A. Stella, G. Lanzani, P. Cheyssac, R. Kofman: Coherent acoustic oscillation in metallic nanoparticles generated with femtosecond optical pulses, Phys. Rev. B 55, R13424 (1997 )
© Copyright 2025 Paperzz