You are on page 1of 10

20 2  Acoustic Field of a 1-D Array Element

and k = ω / c is the wave number. For the real value of k z given in Eq. (2.10), the
solution of Eq. (2.9) represents a plane wave traveling at an angle θ with respect to
the z-axis, where k x = k sin θ , k z = k cos θ . The imaginary value of k z corresponds
to an inhomogeneous wave traveling in the plus or minus x-direction (depending on
the sign of k x ) and decaying exponentially in amplitude in the z-direction. Since the
waves given by Eq. (2.9) are solutions of the Helmholtz equation, we can also form
up a more general solution by simply a superposition of these waves traveling with
different values of k x , i.e. we can let
 +∞
1 (2.11)
p ( x, z , ω ) =
2π ∫ P(k
−∞
x ) exp(ik x x + ik z z )dk x .

This type of solution is called an angular spectrum of plane waves, although as we


have seen it is really a combination of both plane waves and inhomogeneous waves.
We will use this type of solution to represent the waves generated by the element
model of Fig. 2.1. If we let vz ( x, 0, ω ) be the Fourier transform of vz ( x, 0, t ) (on
the plane z = 0) then
 +∞

vz ( x, 0, ω ) = ∫ v ( x, 0, t ) exp(iωt )dt
−∞
z (2.12)

and we see that



vz ( x, 0, ω ) − b < x < b
vz ( x, 0, ω ) =  , (2.13)
0 otherwise

where
 +∞

vz ( x, 0, ω ) = ∫ v ( x, 0, t ) exp(iωt )dt.
−∞
z (2.14)

Note that from the equation of motion (Eq. (2.1)) we have


 ∂p
− = − ρω 2 u z = −iωρvz (2.15)
∂z

so that
 1 ∂p ( x , z , ω )
vz ( x, 0, ω ) = . (2.16)
iωρ ∂z z=0

Thus, from Eq. (2.11) we find


2.1  Single Element Transducer Models (2-D) 21

 1
+∞
ik z
vz ( x, 0, ω ) =
2π ∫ iωρ P(k x ) exp(ik x x)dk x
−∞ (2.17)
+∞
1
=
2π ∫ V (k
−∞
x ) exp(ik x x)dk x,

where
(2.18) ik P(k x )
V (k x ) = z .
iωρ

Equation (2.17) shows that vz ( x, 0, ω ) can be treated as a spatial inverse Fourier


transform of V (k x ) so that from the corresponding forward spatial transform we
have
 +∞

V (k x ) = ∫ v ( x, 0, ω ) exp(−ik x)dx.
−∞
z x (2.19)

Since the velocity on z = 0 is known (Eq. (2.13)) the spatial Fourier transform V (k x )
is also known and we can write the pressure anywhere in the fluid from Eq. (2.11)
and Eq. (2.18) as
 ωρ +∞ V (k x )
p ( x, z , ω ) = ∫ k z exp(ik x x + ik z z )dk x .
2π −∞
(2.20)

Now, Eq. (2.20) is in the form of a spatial inverse Fourier transform of a product of


functions G (k x , z , ω ) and H (k x , ω ), i.e.
 1
+∞
(2.21)
p ( x, z , ω ) =
2π ∫ G(k , z, ω ) H (k , ω ) exp(ik x)dk ,
−∞
x x x x

where

exp(ik z z )
G (k x , z, ω ) = , H (k x , ω ) = ωρV (k x ). (2.22)
kz

But by the convolution theorem [Schmerr] the inverse Fourier transform of a prod-
uct of transformed functions is the convolution of the functions themselves, so that
in this case the convolution theorem gives
 +∞

p ( x, z , ω ) = ∫ h( x ′, ω ) g ( x − x ′, z, ω )dx ′,
−∞
(2.23)
22 2  Acoustic Field of a 1-D Array Element

where
 +∞
1 1 (2.24)
g ( x, z , ω ) =
2π ∫k
−∞ z
exp(ik x x + ik z z )dk x

and

ρω +∞
h ( x, ω ) = ∫ V (k x ) exp(ik x )dk x .
2π −∞
(2.25)

First, examine Eq. (2.25). Since V (k x ) is the spatial Fourier transform of vz ( x, 0, ω ) ,


it follows that
(2.26)
h( x, ω ) = ρω vz ( x, 0, ω ).

Now, consider Eq. (2.24). This is the inverse Fourier transform of an explicit func-
tion and can be shown to be proportional to a Hankel function of zeroth order and
type one [1]. Specifically,


g ( x, z , ω ) =
(
H 0(1) k x 2 + z 2 ), (2.27)
2

where again k = ω /c is the wave number. Placing these results into Eq. (2.21) then
gives
+∞
(2.28) ωρ
p (x, ω ) = ∫ vz ( x ′, 0, ω ) H 0 (kr )dx ′,
(1)

2 −∞

where r = ( x − x ′ ) 2 + z 2 is the distance from a point ( x ′, 0) on the plane z = 0 to


a point x = ( x, z ) in the fluid (see Fig. 2.1). Since the velocity on the plane z = 0 is
given by Eq. (2.13), we find

 ωρ + b (2.29)
p (x, ω ) = ∫ v ( x ′, 0, ω ) H (kr )dx ′
(1)
z 0
2 −b

in terms of the velocity on the face of the element, which is assumed to be known.
Equation (2.29) gives the pressure anywhere in the fluid generated by the motion of
the face of the element so it is a complete ultrasonic model for the waves generated
by a single element radiating into a single fluid medium.
Physically, Eq. (2.29) represents the wave field of the transducer element in
terms of a weighted superposition of cylindrical waves arising from concentrated
sources acting over the length of the transducer. This can be seen more explicitly by
2.2  Far Field Waves 23

Fig. 2.2   Geometry param-


eters for defining the far field
response

assuming the distance r to a point in the fluid is many wavelengths away from the
element so that kr > > 1. Then, since the Hankel function has the asymptotic value
[2]
 2
H 0(1) (u ) = exp [i (u − π / 4) ] (2.30)
πu

for u >> 1, Eq. (2.30) becomes


 +b
k ρc exp( −iπ / 4) 2
p (x, ω ) =
2 ∫
−b
vz ( x ′, 0, ω )
π kr
exp(ikr )dx ′ (2.31)

in terms of a superposition of the cylindrical wave terms exp(ikr )/ r over the


length of the element.

2.2  Far Field Waves

From the law of cosines (see Fig. 2.2) we have


 r = r02 + ( x ′ ) 2 − 2 x ′r0 sin θ . (2.32)

Continuing to keep the high frequency approximation kr >> 1 the far field of the
element is defined as the region far enough from the element so that x ′/r0 << 1 are
valid and we can expand Eq. (2.32) to only first order as
(2.33)
r = r0 − x ′ sin θ.

If we place this approximation into Eq. (2.30) we obtain


 +∞
k ρc 2
p (x, ω ) = exp( −iπ /4) exp(ikr0 ) ∫ vz ( x ′, 0, ω ) exp( −ik sin θ x ′ )dx ′, (2.34)
2 π kr0 −∞
24 2  Acoustic Field of a 1-D Array Element

or equivalently, in terms of the spatial Fourier transform of the velocity field,



k exp(ikr0 )
p (x, ω ) = ρcV (k x ) (2.35)
2π i r0

with k x = k sin θ.
Equation (2.35) shows that in the far field region the element behaves like a con-
centrated source emitting a single cylindrical wave so we could call this region the
cylindrical wave region of the element.
In most cases we will model the motion on the face of an element as if it acted as
a piston source, i.e. as if the element had a spatially uniform velocity over the entire
length of the element:
 v (ω ) − b < x ′ < b
vz ( x ′, 0, ω ) =  0 . (2.36)
0 otherwise

In this case the spatial Fourier transform is


 +b

V (k x ) = ∫ v (ω ) exp(−ik x ′)dx ′
0 x
−b
(2.37)
2v (ω ) sin(k x b) 2v0 (ω ) sin(kb sin θ )
= 0 =
kx k sin θ

and the far field piston element response can be written as



2 sin(kb sin θ ) exp(ikr0 ) (2.38)
p (x, ω ) = ρcv0 (ω ) (kb) .
πi kb sin θ kr0

Equation (2.38) shows that in the far field the piston element response has a direc-
tivity function, Db (θ ), where

sin(kb sin θ ) (2.39)
Db (θ ) = .
kb sin θ

This directivity function is strongly controlled by the non-dimensional wave num-


ber, kb, as shown in Fig. 2.3. [Note: For brevity of notation in later expressions this
kb dependency will be omitted in the argument of Db but it should be implicitly un-
derstood that it is still present in this and in other directivities that will be discussed
in later Chapters.] For a value of kb = 0.314 where the length, 2b, of the element is
one tenth of a wavelength,λ, (Fig. 2.3a), the sound radiation of the element is nearly
uniform in all directions ( −90° ≤ θ ≤ 90° ). At kb = 1.57 (element length = one half
a wavelength) there begins to be some significant changes in directivity with angle
(Fig.  2.3b) but the radiation pattern is still broad. At a value kb = 3.14 (element
2.2  Far Field Waves 25

Fig. 2.3   The directivity function for an element of a length 2b/λ = 0.1, b 2b/λ = 0.5, c 2b/λ = 1.0 ,
and d 2b/λ = 3.0

Table 2.1   Directivities of Element size, 2b/λ


some elements of various
θ −6 dB (degrees)
sizes 1.0 36.9
3.0 11.5
10.0 3.4

length = one wavelength) Fig.  2.3c shows that now most of the radiation is in an
angular region of ±30° about the normal to the element and at kb = 9.42 (element
length = three wavelengths) the sound is confined primarily to a highly directed
beam, with the appearance of small side lobes, as shown in Fig. 2.3d. Most NDE
phased array transducers operate at MHz frequencies and with element sizes that are
larger than one half a wave length so that directivity of the element plays an impor-
tant role in the sound field generated and appears as a part of the overall response
of an array of elements.
It is customary to define the size of the main “lobe” of the far field sound beam
generated by an element in terms of the angle at which the pressure first drops to
one half (− 6 dB) of its value along the z-axis (θ = 0). For the sinc function, sin x/ x,
this one half value occurs at x = 1.8955 so that from Eq. (2.39) we see the − 6 dB
angle is given by
26 2  Acoustic Field of a 1-D Array Element

  λ
θ −6 dB = sin −1 0.6 , (2.40)
 2b 

which always has a root as long as 2b > 0.6λ. Table 2.1 shows the results for cases
(c) and (d) of Fig. 2.3 which agree with the angular patterns shown in Fig. 2.3. Also
shown in Table 2.1 are the results for an element that is ten wave lengths long,
where it can be seen that the directivity becomes quite small. Large, single element
transducers used in NDE applications are normally tens of wavelengths in diameter
so that they are highly directional and generate sound beams that are well collimat-
ed, i.e. most of the sound propagates normal to the face of the transducer. However,
for the smaller elements present in phased array transducers the far field directivity
can vary considerably, depending on the size of the elements.
It is important to know when the far field approximation we have been using in
this section is valid. Recall, in Eq. (2.32) we expanded the radius r to only first order
Eq. (2.33) which led us to the explicit far field results. Let us go back to Eq. (2.32)
and examine when the remaining terms in the expansion are negligible. First, we
rewrite the radius r as
 ( x ′ ) 2 − 2 x ′r0 sin θ
r = r0 1 + , (2.41)
r02

which is in a form that can be expanded to three terms since by the binomial expan-
sion of a square root
  b b2 
1 + b ≈ 1 + − + ... , b < 1. (2.42)
 2 8 

In the case of Eq. (2.41) if we use Eq. (2.42) and keep only quadratic terms at most
in the expansion we find

( x ′ ) 2 cos 2 θ (2.43)
r ≈ r0 − x ′ sin θ + .
2r0

Equation (2.43) shows that in order to keep only the first order term of
Eq. (2.33) in the phase term of Eq. (2.1.31) we must have the complex expo-
nential exp(ik ( x ′ ) 2 cos 2 θ /2r0 ) term near unity, which will only be possible if
k ( x ′ ) 2 cos 2 θ / 2r0 << 1. This condition will certainly be satisfied if we replace x ′
and cos 2 θ by their largest possible values of b and one, respectively, and require

kb 2 π b 2 (2.44)
= << 1.
2r0 λ r0
2.3  Numerical Piston Element Models 27

So the far field conditions should be well satisfied if


(2.45)
π b2
r0 >> .
λ

For large single element transducers radiating waves in 3-D the far field occurs
when the transducer wave field is well approximated by a point source radiating a
single spherical wave. For a circular transducer of radius b the far field of the trans-
ducer is also given by Eq. (2.45) where the quantity N = b 2/λ is called the near
field distance [Schmerr]. The quantity D = π b 2 /λ can be rewritten as D = kb 2/2
which is called the Rayleigh distance [Schmerr-Song]. We see that the far field of a
1-D element also occurs for distances greater than the Rayleigh distance.

2.3  Numerical Piston Element Models

Although we can obtain explicit results for the wave field of an element in the far
field, for other points in the wave field it is not possible to analytically perform the
integrations present in either Eq. (2.29) or Eq. (2.31). However, it is relatively easy
to compute the pressure wave field numerically from Eq. (2.29) as long as z is not
too near the element face ( z = 0) where the Hankel function can become singular.
At or near the element face one needs to perform the integration more carefully, but
generally we are only interested in the radiated wave field at some finite distance
from the element face where such singular behavior is not present. For piston be-
havior we have:

k ρcv0 (ω ) + b (1)
p (x, ω ) =
2 ∫ H 0 (kr )dx ′,
−b
(2.46)

where k = ω /c is the wave number. First, we break up the face of the element into
N equal segments of length Δd = 2b/N. The centroids of these segments will be lo-
cated at the positions

∆d
xn′ = −b + + (n − 1) ∆d (n = 1, 2,...N ) (2.47)
2

or, in normalized form


 xn′ 2 1
xn = = −1 +  n −  . (2.48)
b N 2

If we assume that the segments are small enough that the Hankel function can be
taken as a constant over each segment, then Eq. (2.46) can be written as
28 2  Acoustic Field of a 1-D Array Element

 p (x, ω ) kb N (1)
ρcv0
= ∑ H 0 (kbrn ),
N n =1
(2.49)

where
 (2.50)
rn = ( x /b − xn ) 2 + ( z / b) 2 .

We will call the beam model of Eq. (2.49) the Hankel function model.
Since for large kb values the Hankel function is a highly oscillatory function, to
treat it as a constant over each segment we will typically have to keep the segment
size, Δd, to about a tenth of a wavelength, λ, or less. If we take frequency, f, in MHz,
the wave speed, c, in m/sec, and dimensions of b and Δd in mm, then the criterion
 λ
∆d ≤ (2.51)
10

is equivalent to the requirement that N must satisfy

   20 × 103 fb 
ceil   2b > λ /10 (2.52)
N=  c ,
1 2b ≤ λ /10

where the MATLAB® function ceil() rounds the value of its argument towards in-
finity to the nearest integer. Note that Eq. (2.52) also guarantees that for element
lengths less than a tenth of a wavelength we must have at least one segment ( N = 1).
For b = 6 mm, f = 5 MHz, and c = 1500 m/sec we obtain N = 400. This example shows
that even in this 2-D problem when computing the integral of Eq. (2.46) for element
sizes comparable to large, single element NDE transducers one needs to compute a
relatively large number of terms. For the much smaller element sizes used in phased
arrays the number of terms needed is greatly reduced. The MATLAB® function
rs_2Dv given in Code Listing C.1 (see Appendix C) implements the Hankel func-
tion model. Its calling sequence is

>>p=rs _ 2Dv(b, f, c, e, x, z, Nopt );

This function uses Eq. (2.49) to calculate the normalized pressure field, p ( x, z )/ρcv0
at a point ( x, z ), measured in mm, for a piston element of length 2b mm radiating
waves into a fluid whose wave speed is c (in m/sec) at a frequency, f, given in MHz.
In anticipation that this function will also be used to model a single element in an
array where the center of the element may not be located at x ′ = 0, Eq. (2.50) has
been modified to be
2.3  Numerical Piston Element Models 29

Fig. 2.4   The magnitude of the normalized pressure, p/ρcv0, generated at 5 MHz, c = 1500  m/sec,
for an element where a 2b = .03  mm (2b /λ = 0.1), b 2b = 0.3  mm (2b/λ = 1.0), c 2b = 2.0  mm
(2b/λ = 6.67) , and d 2b = 6.35  mm (2b/λ = 21.2)

 rn = ( x / b − xn − e / b) 2 + ( z / b) 2 , (2.53)

where the offset of the element in the x-direction, e, (in mm) is also specified as an
input parameter of rs_2Dv. The parameter, Nopt, is an optional input argument for
this function. If Nopt is not specified, the number of segments used to calculate the
wave field is given by Eq. (2.52). Otherwise Nopt is used for the number of seg-
ments instead. This gives the user some flexibility in controlling the calculations
but, as we will show shortly, when choosing Nopt there are some limits to how large
the element segments can be.
Figure 2.4 shows the magnitude of the pressure wave field generated by elements of
different sizes. In all these cases the frequency was taken to be 5 MHz and the wave
speed (nominally that of water) was given as 1500 m/sec. The Nopt parameter was
not specified. Figure 2.4a shows the case for an element of total length 2b = 0.03 mm,
which corresponds to 2b /λ = 0.1. For this case in the far field we saw the directivity
of the element was uniform in all directions and this behavior is also true for much of

You might also like