21.06.2014 Views

A General Purpose Fiber-Optic Hydrophone Made of Castable Epoxy

A General Purpose Fiber-Optic Hydrophone Made of Castable Epoxy

A General Purpose Fiber-Optic Hydrophone Made of Castable Epoxy

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

A <strong>General</strong> <strong>Purpose</strong> <strong>Fiber</strong>-<strong>Optic</strong> <strong>Hydrophone</strong> <strong>Made</strong> <strong>of</strong> <strong>Castable</strong> <strong>Epoxy</strong><br />

Steven L. Garrett, David A. Brown, B. L. Beaton(a), K. Wetterskog(a), and J. Serocki(a)<br />

Naval Postgraduate School, Physics Department, Code PH/Gx<br />

Monterey, CA 93943<br />

ABSTRACT<br />

A fiber-optic, interferometric, flexural disk hydrophone cast from an epoxy resin is described. This<br />

hydrophone is designed in the shape <strong>of</strong> an enclosed, hollow cylinder with pairs <strong>of</strong> flat, spiral wound coils<br />

<strong>of</strong> optical fiber embedded in each sensing plate. An all-fiber Michelson interferometer is used to detect the<br />

optical phase shift which results from pressure induced strains in the optical fiber. The sensing coils are<br />

positioned in the plates in a manner to enhance the acoustic response and provide cancelation <strong>of</strong><br />

acceleration induced signals. An epoxy resin was chosen for its relatively high tensile strength, its low<br />

Young's modulus, and its ability to cure at room temperature. The acoustic sensitivity <strong>of</strong> this sensor in<br />

both air and water was measured to be 0.277 0.005 rad/Pa (-131.2 dB re radfliPa) which corresponds to<br />

a normalized sensitivity 4/p = a1n4/ap<br />

= -297 dB re 1 iPa1 below 1 .0 kHz. This measured result is in<br />

excellent agreement with simple elastic theory and the measured epoxy elastic constants. The normalized<br />

acceleration sensitivity is 3/4aa = ln/aa -163 dB re g1. The acceleration-to-acoustic sensitivity ratio<br />

(figure-<strong>of</strong>-merit) <strong>of</strong> -134 dB re g4tPa is the largest reported to date for any fiber-optic hydrophone.<br />

1. INTRODUCTION<br />

<strong>Optic</strong>al fibers alone are neither intrinsically sensitive nor selective, hence it can be difficult to design a<br />

fiber-optic hydrophone which is sensitive to acoustic pressure while rejecting variations in ambient<br />

pressure and temperature, platform vibration, flow noise, or other environmental quantities1. Most <strong>of</strong> the<br />

published hydrophone research has involved the use <strong>of</strong> one coating to increase sensitivity to the measurand<br />

<strong>of</strong> interest for one on the interferometer legs and a different coating or decoupling scheme to desensitize the<br />

other interferometer leg which acts as a reference2'3. We feel that a better way to overcome the selectivity<br />

problem and to increase sensitivity to the measurand <strong>of</strong> interest is to design fiber-optic interferometric<br />

sensors in which both legs <strong>of</strong> the interferometer are active and respond to the measurand field with strains<br />

(i.e., changes in optical path-length) <strong>of</strong> opposite sign. Since the interferometer generates a signal in<br />

response only to differences in the optical path lengths in the two arms, the interferometer automatically<br />

provides common-mode rejection <strong>of</strong> the unwanted influences. The fact that the optical path-length in both<br />

arms change in the opposite direction also doubles the sensitivity (+6 dB) <strong>of</strong> the hydrophone over that <strong>of</strong><br />

one having an identical construction (i.e., same materials, fiber length, etc.) but employing an "ideal"<br />

reference arm. The success <strong>of</strong> this strategy has been documented in the open literature by both the Naval<br />

Postgraduate Schoo1418 and Litton Guidance and Control19'2° and has produced the first fiber-optic<br />

sensor whose detection threshold was limited by its own temperature21.<br />

SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990) / 13


2. DUAL FLEXURAL PLATE HYDROPHONE ELEMENT DESIGN<br />

2 . 1 Pressure and Acceleration Response<br />

When .a thin circular plate is subject to a pressure differential across its surface, it will deform and<br />

therefore experience a surface strain which is compressive on one side and expansive on the other. This<br />

fact has been used previously to develop a hydrophone element using piezoelectric polymer films as the<br />

strain transducers22'23. We have applied this strategy to fiber-optic hydrophone designs by bonding flat<br />

spiral coils <strong>of</strong> optical fiber to both surfaces12 14,17, 1 8 jf the two coils act as the two legs <strong>of</strong> an<br />

interferometer, the plate becomes the transduction mechanism in a push-pull fiber-optic hydrophone.<br />

A single-plate hydrophone will also sense accelerations in the direction normal to the plate surface<br />

because the mass <strong>of</strong> the plate is not zero. Such an acceleration would produce a deflection which is equal<br />

to that produced by a pressure differential given by the product <strong>of</strong> the plate mass and its acceleration<br />

divided by the surface area <strong>of</strong> the plate. The acceleration sensitivity can be reduced by using two identical<br />

plates and four coils as shown schematically in Figures 1 and 2. When the element is subjected to an<br />

excess external pressure, both plates are displaced inward toward the air gap. This deformation causes the<br />

length <strong>of</strong>both Al and A2 to decrease while both Bi and B2 increase producing a "push-pull" enhancement<br />

<strong>of</strong> the optical path length change due to the applied pressure excess. The signs <strong>of</strong> these changes would<br />

reverse for a pressure reduction. This doubles the element's sensitivity to pressure over a single plate<br />

hydrophone <strong>of</strong> identical design.<br />

- If the entire sensor element is accelerated in a direction which is normal to the plate surface, then the<br />

two plates move in unison and the length <strong>of</strong> Al and B2 would increase while the length <strong>of</strong> A2 and Bl<br />

would decrease. The signs <strong>of</strong> these changes would reverse for a reversal in the direction <strong>of</strong> acceleration.<br />

Since the coils in each arm <strong>of</strong> the interferometer ideally experience no net change in length in either arm,<br />

AA1 + L\A2 = AB1 + EB2 = 0, the element is, in principal, immune to accelerations. If the plates, and the<br />

location <strong>of</strong> the fiber-optic coils in each plate, are not identical, the cancelation will not be optimized, so<br />

careful attention has been paid in the fabrication process to make these elements as symmetric as possible<br />

about both the air gap and the rotation axis.<br />

Al<br />

A2<br />

3 dB<br />

Bl 52<br />

Fig. 1. <strong>Fiber</strong>-optic Michelson interferometer. Each sensing leg consists <strong>of</strong> two flat spiral coils 4.50m in<br />

length and 4.0 cm in diameter. The two coils labeled Al and A2 are placed on the outer surfaces <strong>of</strong> the<br />

sensing plates and the two coils B 1 and B2 are placed on the inner surfaces which face the air gap.<br />

14 / SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990)


ATMOSPHERIC PRESSURE<br />

Sensing Plate 1<br />

(a)<br />

<strong>Fiber</strong> Coils<br />

:: r5::::e: Air Gap<br />

AirGap<br />

Sensing Plate 2<br />

(a)<br />

PRESSURE RESPONSE<br />

1<br />

(b)<br />

Sensing Plates<br />

(b)<br />

TEST DEPTH<br />

(C)<br />

(C)<br />

Fig. 2. (Left) Schematic representation <strong>of</strong> the four-coil, dual-plate hydrophone element showing a crosssectional<br />

view (a) with the dimensions <strong>of</strong> the plates, air gap, and stem (coupler housing). The location <strong>of</strong><br />

the fused evanescent wave directional coupler and the fiber coils is shown in cross-section in (b) and in<br />

plan view in (c).<br />

Fig. 3. (Right) Static pressure response <strong>of</strong> a dual-plate flexural disk hydrophone. Cross-sectional views<br />

show the position <strong>of</strong> the plates and the air gap at (a) atmospheric pressure, (b) operating pressure, and (c)<br />

maximum operating depth where the plates touch.<br />

2.2 Dynamic Flexural Plate Theory<br />

We have previously constructed and tested several flexural plate hydrophone designs using 6061-T6<br />

aluminium and have found that their measured sensitivity and their theoretical sensitivity, based on the<br />

elastic properties <strong>of</strong> the plate material and the boundary conditions, are in excellent agreement for both<br />

hydrostatic and acoustic pressure fie1ds1214'17. The sensitivity, M = (a/P), <strong>of</strong> the dual-plate, fourcoil<br />

hydrophone element, assuming that the perimeter <strong>of</strong> the plate is rigidly clamped, can be expressed in<br />

terms <strong>of</strong> the length <strong>of</strong> the optical fiber in one leg, L, the elastic properties <strong>of</strong> the plate material, and the plate<br />

and fiber coils' physical dimensions.<br />

SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990) / 15


M=a<br />

8icnL<br />

2<br />

- b2 - c2)<br />

(1)<br />

Here 4 represents the optical pathlength in radians and P is the applied excess acoustic pressure. E and<br />

are the Young's modulus and Poisson's ratio <strong>of</strong> the plate material, 2h is the plate thickness, and a,b, and<br />

c, are the radius <strong>of</strong> the plate and the maximum and minimum spiral coil radii respectively. The wavelength<br />

<strong>of</strong> light in vacuum is , and n is the effective index <strong>of</strong> refraction <strong>of</strong> the optical fiber core. The derivation <strong>of</strong><br />

this result and similar results for simply supported plates with bonded optical fiber spirals have been<br />

published elsewhere13'16'17.<br />

It is conventional to normalize this sensitivity to provide a comparison with other fiber-optic<br />

hydrophones that may use different mechanical designs, fiber lengths, optical wavelengths, and glasses<br />

with different indices <strong>of</strong> refraction. The normalized sensitivity, M = = is<br />

determined by dividing the acoustic sensitivity by the optical path length (in radians <strong>of</strong> phase) <strong>of</strong> one leg <strong>of</strong><br />

the sensor. This convention has been used because most early interferometric sensors use one leg as a<br />

reference leg. The sensitivity can be expressed in a particularly simple form for two plates totally wrapped<br />

on each side with optical fiber and having ideal clamped boundary conditions as<br />

M=O.52—<br />

(2)<br />

Hence, the product <strong>of</strong> the sensitivity and the square <strong>of</strong> the fundamental plate resonance frequency, f02, is a<br />

constant which depends only upon the wavenumber <strong>of</strong> the light in the fiber, kg 2ltflIX, the density <strong>of</strong> the<br />

plate material, p, and the outside diameter <strong>of</strong> the optical fiber, D. For the simply supported case, the prefactor<br />

in (2) is equal to 0. 125 (5 + c)/(1 + ). When 0.27, the pre-factors for both cases are identical.<br />

The fundamental resonance frequency <strong>of</strong> the plate is given by<br />

f0=A2J1—4F E (3)<br />

ita2 V 12(1-&)p<br />

where A2 = 10.2 for a clamped plate. For a simply supported plate, A2 4.9, and is a very weak function<br />

<strong>of</strong> Poisson's ratio.<br />

By Newton's Second Law, the acceleration, a, <strong>of</strong> a plate normal to its surface produces a force which<br />

is the product <strong>of</strong> the mass <strong>of</strong> the plate and its acceleration. That force, divided by the area <strong>of</strong> the plate,<br />

produces an equivalent acceleration induced pressure. One can use this "equivalence" to write an<br />

expression for the acceleration sensitivity, Ma, for a plate <strong>of</strong> mass m = 2itpha2.<br />

M=2.=2Mhp<br />

(4)<br />

16 / SP/E Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors Vii (1990)


The reader is reminded that M is the acoustic sensitivity as defined in equation (1). This result is useful in<br />

the determination <strong>of</strong> how successful the dual-plate design has been in canceling the acceleration induced<br />

forces. An ideal acceleration canceling hydrophone would have a measured Macc which would be<br />

identically zero. This, however, is very difficult to achieve since the physical dimensions, the plate<br />

boundary condition, the plate mass, and the length <strong>of</strong> fiber in each coil and their distance from the plate's<br />

neutral axis would all have to be identical for both sensing plates.<br />

2 . 2 Static Flexural Plate Theory<br />

In any hydrophone application one has also to consider the response to changes in operating depth.<br />

Because this design is "air backed", it is important to limit the deformation <strong>of</strong> the plates for two reasons.<br />

The first is simply structural. If the plate is constructed from a material with a yield strength, Tyjeld, the<br />

maximum operating pressure, m' must be limited to4,<br />

max 16h2 Tyjeld (5)<br />

3a<br />

The second reason the plate deformation must be limited is unique to fiber-optic hydrophones and is due to<br />

the fact that the maximum strain which can be tolerated by the typical glass fiber with breakage17 is<br />

approximately 0.5%.<br />

If one chooses not to employ pressure compensation techniques25 to equilibrate the static pressure<br />

inside and outside the hydrophone element due to the added complexity or degradation <strong>of</strong> the low<br />

frequency response, 1max a function <strong>of</strong> the plate dimensions and material and is inversely proportional<br />

to the hydrophone sensitivity. Equations (1) and (5) can be combined to produce an approximate<br />

expression for the normalized sensitivity <strong>of</strong> a hydrophone element <strong>of</strong> this flexural plate design.<br />

I M = = aln Tyjeld<br />

n<br />

4E max<br />

ap , A.<br />

ap<br />

(6)<br />

Approximations involved in deriving equation (6) include assuming a clamped boundary condition, letting<br />

a = b >> c, and setting factors such as (1 - 2), which are close to one, equal to one. Equation (6)<br />

suggests that the optimum material for this style <strong>of</strong> hydrophone element would be one which has the<br />

greatest ratio <strong>of</strong> yield strength to Young's modulus. Plastics tend to be about five times better than<br />

aluminum based on that criterion22 although the temperature dependance <strong>of</strong> their Young's modulus is<br />

significantly larger2 and the resulting resonance frequency and bandwidth is also lowered.<br />

By Hooke's Law, the maximum stresses can be related to the maximum strains and hence,17 the<br />

maximum strain in the optical fiber occuring at max' can be expressed using the same<br />

approximations as used in equation (6), for a clamped disk as<br />

(AL) max<br />

Tyjeld<br />

SPIE Vol 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990) / 17


Since the normalized sensitivity is proportional to AL/L, the sensitivity <strong>of</strong> a hydrophone element based on<br />

the integrated strain on the surfaces <strong>of</strong> a flexural plate will also be proportional to the ratio <strong>of</strong> yield strength<br />

to Young's modulus.<br />

It is possible to use the size <strong>of</strong> the air gap in the dual-plate hydrophone element to control the maximum<br />

strain experienced by the glass fiber and to provide support at depths which are greater than those given be<br />

equation (5)23. This is accomplished, as shown in Figure 3, by choosing the plate spacing provided by<br />

the air gap to be twice the displacement <strong>of</strong> a single plate at the maximum operating pressure. The<br />

displacement <strong>of</strong> the center <strong>of</strong> a thin clamped plate, z, is given26 as<br />

3(1 -<br />

zc=<br />

128Eh3<br />

(8)<br />

With the plates in contact, the hydrophone element will be protected from catastrophic failure at the<br />

expense <strong>of</strong> becoming acoustically insensitive until the pressure is once again reduced below 1max<br />

3.1 Material Selection<br />

3. HYDROPHONE ELEMENT CONSTRUCTION<br />

A castable epoxy was chosen as the capsule material so that the fiber coils could be cast directly into<br />

the sensing plates <strong>of</strong> the hydrophone. This also facilitated the fabrication <strong>of</strong> plates which were <strong>of</strong> uniform<br />

thickness. Casting the sensor capsule as a single unit also ensures that the boundary conditions at the wall<br />

were well matched and reproducible for both plates.<br />

TABLE 1. Measured Material Properties <strong>of</strong> E-CAST F-28<br />

Property<br />

Mass density<br />

Dynamic Young's modulus (E)<br />

Dynamic shear modulus (G)<br />

Thermal coefficient <strong>of</strong> dynamic elasticity (lnE/aT)<br />

Thermal coefficient <strong>of</strong> dynamic elasticity ()lnG/aT)<br />

Static Young's modulus (Estat)<br />

Tensile strength (Tyjeld)<br />

Poisson's ration ()<br />

Strength-to-modulusratio (TyjeidE)<br />

Value<br />

1.15<br />

3.4 0.2<br />

1.15 0.09<br />

0.41 0.2<br />

0.39 0.2<br />

3.0 0.2<br />

6.8 0.4<br />

0.39<br />

0.02<br />

Units<br />

gm/cm3<br />

iO Pa<br />

10 Pa<br />

%/°C<br />

%/°C<br />

10 Pa<br />

10 Pa<br />

The process which led to the selection <strong>of</strong> E-CAST F-28 with 215 hardner27 as the hydrophone<br />

element body and plate material involved the testing <strong>of</strong> seventeen different room temperature castable<br />

epoxy compounds and composites to determine their static Young's modulus, yield strength, and dynamic<br />

shear and Young's moduli as a function <strong>of</strong> temperature in the range from 0°C to 25°C and for frequencies<br />

18 / SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990)


on the order <strong>of</strong> one kilohertz. The results <strong>of</strong> these measurements, on all seventeen materials, will be<br />

reported separately28 as well as the method29 and apparatus3° which was developed to automate the<br />

measurements <strong>of</strong> the dynamic shear and Young's moduli as a function <strong>of</strong> temperature. The static Young's<br />

modulus and tensile strength were determined by conventional techniques31. The measured properties <strong>of</strong><br />

the E-CAST F-28 are listed in Table 1. This was the only epoxy material used in the construction <strong>of</strong> the<br />

capsule so in all subsequent discussion the term epoxy will be understood to refer to E-CAST F-28.<br />

3 . 2 Resonance Frequency<br />

The gravest resonance frequency <strong>of</strong> the capsule sets an upper limit on the hydrophone bandwidth and<br />

also provides insight into the actual boundary conditions experienced by the plates. The theoretical<br />

resonance frequency <strong>of</strong> the sensing plates can be calculated from equation (3). Assuming ideally clamped<br />

plate boundary conditions, a resonance frequency <strong>of</strong> 1 1 .0 kHz was predicted for each plate. If the<br />

boundary was simply supported, rather than clamped, the resonance frequency would be 5.3 kHz. The<br />

actual resonance frequencies for both plates were measured by tapping on the plates and recording the<br />

resulting free decay response with a microphone and digital storage oscilloscope. Based on the period <strong>of</strong><br />

twenty cycles, the resonance frequency <strong>of</strong> the top and bottom plates was determined to be 7.79 0.03<br />

kHz and 7.77 0.01 kHz respectively. Since the measured resonance frequencies fall between the two<br />

ideal cases, the boundary conditions were considered to be a compromise between ideally clamped and<br />

simply supported.<br />

To further investigate the modes <strong>of</strong> vibration, a finite element model <strong>of</strong> the capsule was generated.<br />

That model predicted a resonance frequency <strong>of</strong> 7. 14 kHz which was within eight percent <strong>of</strong> the measured<br />

result. That small difference may be due to the perturbations induced by fiber coils and the attachment <strong>of</strong><br />

the stem which housed the coupler, neither <strong>of</strong> which were included in the model. The modal shape is<br />

shown in the upper portion <strong>of</strong> Figure 4. It is interesting to note that a slight deformation <strong>of</strong> the outer wall<br />

occurs and hence the boundary condition is truly elastic33. The resonant mode for an acceleration induced<br />

response occurs at 8.95 kHz and is shown in the lower portion <strong>of</strong> Figure 5. If the plates are identical, this<br />

mode produces no net change in the interferometer's optical path length and therefore, it can be employed<br />

to evaluate the effectiveness <strong>of</strong> the acceleration canceling design.<br />

3 . 3 Interferometer Fabrication<br />

The objective <strong>of</strong> this sensor design was to provide a means <strong>of</strong> fabrication that was relatively simple,<br />

yet integrates delicate fiber optic components into a single, rugged capsule that is suitable for sustained<br />

underwater use. The first step in the fabrication process was the construction <strong>of</strong> the all-fiber, Michelson<br />

interferometer.<br />

The flat spiral optical fiber coils were wound by hand between two lucite plates separated by a washer<br />

with a thickness <strong>of</strong> approximately 300 .tm. This allowed the optical fiber which had an acrylate coating<br />

with an outer diameter <strong>of</strong> 250 m to pass between the plates but prevented the individual turns from<br />

sliding over each other during the winding process. Each coil consisted <strong>of</strong> 4.50 meters <strong>of</strong> fiber with<br />

approximately one centimeter <strong>of</strong> fiber between the coil pairs. The coils were bonded with thin strips <strong>of</strong><br />

rubber adhesive before they were removed from the winding jig to prevent unwinding prior to their<br />

attachment to the sensing plates.<br />

SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990) / 19


After the coils were wound, the input lead <strong>of</strong> the coupler34 was spliced to a pigtailed Hitachi HL83 12G<br />

laser diode35. A Litton Model 87B <strong>Fiber</strong> <strong>Optic</strong> Sensor Demodulator36 was used to power to the laser<br />

diode, modulate the laser output wavelength by modulating the diode current, and monitor the output <strong>of</strong><br />

the interferometer. A 20 kHz current modulation was applied to the laser diode. By monitoring the optical<br />

output <strong>of</strong> the interferometer, the fiber ends were cleaved to achieve an optimal path length difference<br />

between the two interferometer legs for demodulation purposes and to maximize the modulation depth<br />

(i.e., fringe visibility). The result was a difference in length <strong>of</strong> 2.4 cm. This was small enough to ensure<br />

operation within the coherence length <strong>of</strong> the laser diode and large enough to avoid the use <strong>of</strong> an<br />

unacceptably high laser diode current modulation amplitude. At this point the cleaved ends were protected<br />

by using an adhesive to bond plastic sleeves over the ends <strong>of</strong> the fiber.<br />

<strong>Fiber</strong> leads 7<br />

from couplerf<br />

L<br />

Wax disk<br />

<strong>Fiber</strong> coils<br />

(a)<br />

(a)<br />

Inner Core<br />

Thin epoxy layer<br />

Outer<br />

Capsule<br />

Shell<br />

Walls <strong>of</strong> sensor<br />

(b)<br />

Fig. 4. (Left) Finite element model <strong>of</strong> capsule mechanical resonances. a) Fundamental pressure<br />

response resonant mode. b) Fundamental acceleration response resonant mode.<br />

Fig. 5. (Right) Sensor inner core construction. a) Exploded view <strong>of</strong> "sandwich. b) Inner core within<br />

completed epoxy capsule.<br />

20 / SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors V/Il (1990)


3 .4 Capsule Fabrication<br />

The task <strong>of</strong> packaging the four interconnected fiber-optic coils and their coupler in a compact unit with<br />

an air gap <strong>of</strong> controlled size is not trivial. Since the two sensing coils are connected by the delicate fibers,<br />

the capsule can not be easily molded in two sections and later bonded into a single unit. The solution to<br />

this problem was to pour the capsule in one step as a single unit and create the air gap with parafin' wax<br />

using the "lost wax" technique.<br />

An inner core was constructed by making the "sandwich" shown schematically in Figure 5. The inner<br />

spacing disks were used to maintain an equal spacing between the coils in each plate. The circular wax<br />

disk was formed by pouring melted paraffin in a mold. Two 2-56x 1/4" machine screws were set into the<br />

wax disk to provide "feet" for the inner core shown in Figure 6. These screws provided for alignment <strong>of</strong><br />

the sensor when standing upright in the outer shell mold and also provided two ducts for the removal <strong>of</strong><br />

the wax after the assembled capsule had cured.<br />

Fig. 6. Photograph <strong>of</strong> an assembled inner core showing coupler at the left and machine screw "feet" at the<br />

right.<br />

Fig. 7. Photograph <strong>of</strong> the completed capsule.<br />

SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors V/il (1990) / 21


The inner core was placed in a lucite mold which could then be filled with epoxy to form the capsule.<br />

After the epoxy had cured, the capsule was removed from the mold and the two machine screws were<br />

removed from the capsule. The capsule was then placed in a vacuum oven and heated for approximately<br />

twelve hours at 500 C to remove the wax and post-cure the epoxy. Before encapsulation, both screw holes<br />

were filled with epoxy, but at this stage only one hole was filled so that the other hole could be used to<br />

connect the capsule to a vacuum rig for static pressure sensitivity calibration. A photograph <strong>of</strong> the<br />

completed capsule is shown as Figure 7.<br />

4. HYDROPHONE CALIBRATION AND COMPARISON TO THEORY<br />

The results reported in this section were obtained by the comparison technique39 in air and in water<br />

near atmospheric pressure in the calibrators which were small compared to the acoustic wavelengths <strong>of</strong><br />

interest since the hydrophone was compact'0 (i.e., small compared to the wavelengths <strong>of</strong> interest) and its<br />

response is expected to be omnidirectional The hydrophone was tested in water over the range <strong>of</strong><br />

temperatures from 7.5 °C to 45 °C, at ambient pressure.<br />

The acoustical and vibration sensitivities were measured by a fixed phase shift techniquel8 which<br />

involved the adjustment <strong>of</strong> a sinusoidal excitation (acoustic pressure or acceleration) until the output <strong>of</strong> the<br />

interferometer, observed on an oscilloscope, corresponded to an identifiable amount <strong>of</strong> optical phase shift<br />

(integer multiples <strong>of</strong> 2it radians per half-cycle <strong>of</strong> excitation). This technique can provide results which<br />

were accurate to better than and agreed with the Bessel function zero-crossing technique4'5 also to<br />

better than 1 %. The results reported here are therefore demodulator independent. The normalized<br />

sensitivity, M, was obtained from the measured sensitivities by subtracting 20 log10 (2nkL) from the<br />

measured sensitivities expressed in decibels. The factor <strong>of</strong> two appears due to the "double pass" in a<br />

Michelson interferometer. For this hydrophone, n = 1.48, A = 830 nm, and L = 9.0 m, which<br />

corresponds to a normalization factor is 166 dB.<br />

4 . 1 Comparison Calibration Standards<br />

Measurements <strong>of</strong> acoustic pressure sensitivity in air were made between 60 Hz and 900Hz using a 1-<br />

inch laboratory standard condenser microphone (Brüel & Kjer Model 4132, S/N 172894) as a transfer<br />

standard. The calibration <strong>of</strong> this microphone with its associated pre-amplifier (Brüel & Kjer Model<br />

2660), cable, and power supply (BrUel & Kjer Model 2807) was checked using a pistonphone (Brüel &<br />

Kjer Model 4220) for acoustic calibration prior to and immediately after use. This calibration was verified<br />

with a high precision reciprocity calibration <strong>of</strong> the 1-inch microphone using the BrUel & Kjer Type 4143<br />

Reciprocity Calibration Apparatus and two additional BrUel & Kjar 1-inch reference microphones. The<br />

measured sensitivity was 48.3 mV/Pa.<br />

Measurements <strong>of</strong> acceleration sensitivity were made between 40 Hz and 1 kHz using a piezoelectric<br />

accelerometer (BrUel & Kjar Type 4382) followed by a voltage pre-amplifier (Ithaco Model 1201). The<br />

calibration <strong>of</strong> the accelerometer, cable, and amplifier were checked 159 Hz using a calibration exciter<br />

(Brüel & Kjar Type 4294) which was itself check by comparison to the "chatter" method based on the<br />

magnitude if the Earth's gravitational acceleration41. The two calibrations were found to agree to within<br />

2%.<br />

22 / SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990)


Acoustic calibrations in water were made between 14 Hz and 300 Hz in a water-filled compliant tube<br />

calibrator2. A Brüel & Kjar Type 8103 hydrophone followed by an Ithaco Model 1201 pre-amplifier<br />

served as the reference hydrophone. The hydrophone, cable, and pre-amplifier were calibrated in air using<br />

a BrUel & Kjer Type 4223 Calibrator.<br />

4 . 2 Acoustic Sensitivity<br />

The acoustical sensitivity <strong>of</strong> the hydrophone capsule measured in water and air near room temperature<br />

are summarized in Figure 8, and yield an average sensitivity from 14 Hz to 900 Hz <strong>of</strong> 0.277 0.005<br />

radians/Pa (±0.16 dB), for peak acoustic pressure amplitudes in the range <strong>of</strong> 1 1 to 95 Pa in air and 45 to<br />

480 Pa in water. As expected, the sensitivities measured in air and water are identical. The corresponding<br />

normalized acoustic sensitivity is d/dp = a1n4/ap = -297 dB re 1 pPa1. The predicted sensitivity based<br />

on equation (1) is 0.263 radians per Pascal, where we have set n = 1.48, L = 4.5 m, = 0.39, a = 2.0<br />

cm, b = 0.82 cm, c = 1.96 cm, E = 3.4 x iO Pa, h = t/2 = 0.25 cm, and = 830 nm.<br />

This excellent agreement between measurement and theory is not fortuitous! Although the assumption<br />

underlying equation (1) is that <strong>of</strong> an ideally clamped boundary condition, the prefactor for equation (2),<br />

which expresses the sensitivity in terms <strong>of</strong> the fiber properties and the flexural plate resonance frequency,<br />

is insensitive to the specific boundary condition. For the perfectly clamped case it is 0.52 and for the<br />

simply supported case it is 0.48. If we take the intermediate case prefactor to be 0.5 and include a 10%<br />

increase in the fiber diameter (250 tim) to allow for wrapping and adhesive, the measured resonance<br />

frequency (7.8 kHz) yields a boundary condition independent sensitivity <strong>of</strong> 0.29 radians per Pascal.<br />

4 . 3 Static Pressure Sensitivity<br />

As shown in Table 1, the static Young's modulus <strong>of</strong> the plate material was measured to be 12% lower<br />

than the dynamic Young's modulus which suggests, based on equation (1), that the static pressure<br />

sensitivity should be 1 2% greater than the acoustic sensitivity. This was measured by evacuating the air<br />

gap within capsule through one <strong>of</strong> the wax drainage ducts and recording the optical fringes on a strip chart<br />

recorder while the pressure was slowly equilibrated. The static sensitivity, fl radians per Pascal<br />

can be expressed as<br />

Mstatic = 1.856 x i0 -u--- (9)<br />

AP<br />

where N is the number <strong>of</strong> interferometric fringes produced while the pressure is changed by an amount,<br />

Al?, expressed in inches <strong>of</strong> mercury. The prefactor in (9) arises from the fact that a "fringe" corresponds<br />

to 2it radians <strong>of</strong> optical phase shift and 1013 mbar equals 29.92 inches Hg.<br />

The capsule was evacuated to 0. 1 in Hg and was allowed to leak at an initial rate <strong>of</strong> 0.6 in Hg per<br />

minute. Three thousand two hundred ninety-two fringes were counted when the pressure reached 20.0<br />

0.2 in Hg. This produced a static sensitivity <strong>of</strong> 0.305 radians per Pascal which is about 10% higher than<br />

the acoustic measurements. This result very close to what would be expected since the measured<br />

differences between the static and dynamic Young's moduli differ by 12%.<br />

SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990) / 23


C. 5<br />

C-.<br />

Ct<br />

D.. —<br />

P<br />

><br />

C.3—<br />

am 4t am ,v,. - -<br />

cL 'p w'<br />

Cs)<br />

a)<br />

Li)<br />

0<br />

0.2<br />

U)<br />

C<br />

0<br />

100<br />

Freouenc, hertz<br />

DOD<br />

Fig. 8 Measured acoustic sensitivity <strong>of</strong> the hydrophone capsule in air and water. The solid line represents<br />

a best fit to the measured data and has a value <strong>of</strong> 0.277 0.005 rad/Pa. The water measurements are<br />

represented by the filled circles and were made at 23° C. The air measurements are represented by the<br />

open diamonds and were made at 22° C.<br />

4 . 4 Temperature Coefficient <strong>of</strong> Acoustic Sensitivity<br />

The hydrophone capsule was calibrated in water over the temperature range from 7.5° C to 45° C<br />

(45.5° F to 1 1 3° F) and the sensitivity was found to change at the rate <strong>of</strong> +1 .1% /°C . That is, the<br />

sensitivity <strong>of</strong> the capsule increases with increasing temperature (+0. 10 dB/°C). This result includes the<br />

correction for the temperature coefficient <strong>of</strong> the sensitivity <strong>of</strong> the Bruel and Kjer Type 8103 which is given<br />

by the manufacturer as -0.03 dB/°C.<br />

Based on the modulus measurements <strong>of</strong> the epoxy18'23 shown in Table 1, one would expect a<br />

temperature coefficient <strong>of</strong> acoustic sensitivity that was equal to the measured temperature coefficient <strong>of</strong> the<br />

Young's modulus which was only +0.4 %/°C. In an attempt to check this result, the temperature<br />

coefficient <strong>of</strong> the elastic moduli were remeasured using a variation <strong>of</strong> technique and reaffirmed the original<br />

results.<br />

At the present time, the best available explanation for this discrepency is that the temperature<br />

dependance <strong>of</strong> the Young's modulus <strong>of</strong> the epoxy also effected the boundary conditions (and the<br />

resonance frequency) and therefore had a greater effect than would be predicted from consideration <strong>of</strong><br />

equations (1) or (2). The finite element model in Figure 5a tends to reinforce this hypothesis since the<br />

elastic response <strong>of</strong> the junction <strong>of</strong> the two plates is clearly also flexing. Further studies will be attempted<br />

to resolve and quantify this discrepancy.<br />

24 / SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors V/li (1990)


4 . 5 Acceleration Sensitivity and Cancelation Figure-<strong>of</strong>-Merit<br />

The hydrophone capsule was mounted on the surface <strong>of</strong> an APS 120S shaker table using a support<br />

ring with two BrUel and Kjar Type 4382 piezoelectric accelerometers as shown in Figure 9. The use <strong>of</strong><br />

two accelerometers ensured that the acceleration experienced by the hydrophone capsule mounted between<br />

the two accelerometers was uniform. The results <strong>of</strong> these measurements for accelerations in the direction<br />

normal to the sensor plate surfaces is shown in Figure 10. The filled dots represent the measurement<br />

when the capsule stem was supported and show an average acceleration sensitivity between 70 Hz and 900<br />

Hz <strong>of</strong> 1.4 0.6 rad/g (-163 dB re 1 g1), where g is the acceleration due to gravity (9.8 m/sec2). For<br />

comparison, the predicted single-plate, dual-coil acceleration sensitivity, based on equation (4) is shown as<br />

a solid line at 7.4 rad/g. The X's represent the measured acceleration sensitivity <strong>of</strong> the capsule when the<br />

stem containing the coupler and about 1 .5 cm <strong>of</strong> optical fiber is not supported on the shaker table. It is<br />

clear that the stem has a cantilever flexural resonance at about 300 Hz which would significantly degrade<br />

the acceleration response if neglected. The transverse acceleration sensitivity was measured by mounting<br />

the capsule on the shaker table with the stem pointing upward. The average acceleration sensitivity in this<br />

orientation, between 250 Hz and 1 kHz was determined to be 1 .8 0.7 rad/g (-161 dB re 1 g1), again<br />

with the stem unsupported.<br />

To estimate the success <strong>of</strong> the acceleration canceling design, one can form the ratio <strong>of</strong> the acceleration<br />

output when the two plates are "wired" to enhance acceleration sensitivity44 (and cancel pressure) to the<br />

output when the plates are in their acceleration canceling configuration. If both plates were used to<br />

measure acceleration the sensitivity would be twice the single plate sensitivity or 14.8 rad/g. Since the<br />

average measured acceleration sensitivity was 1.4 rad/g, the design can be said to provide slightly better<br />

than 20 dB <strong>of</strong> acceleration rejection independent <strong>of</strong> any additional vibration isolating mounting.<br />

Figure 9. Photograph <strong>of</strong> the hydrophone capsule mounted on the shaker table with two piezoelectric<br />

accelerometers.<br />

SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990) / 25


ic.D<br />

><br />

.-)<br />

C<br />

Cr)<br />

C<br />

0<br />

.-)<br />

0<br />

a)<br />

a)<br />

0<br />

e.<br />

6.0<br />

4.0<br />

2.0<br />

liii ZI1II 1ZI ZI 1.<br />

x<br />

xXZ)<br />

.<br />

--;-;-<br />

T<br />

0.0<br />

40 ido<br />

.<br />

><br />

Figure 10. Axial acceleration sensitivity. The filled dots represent the measurement when the capsule<br />

stem was supported and show an average acceleration sensitivity between 70 Hz and 900 Hz <strong>of</strong> 1.4 0.6<br />

rad/g, represented by the dashed line. For comparison, the predicted single-plate, dual-coil acceleration<br />

sensitivity, based on equation (4) is shown as a solid line at 7.4 rad/g. The X's represent the measured<br />

acceleration sensitivity <strong>of</strong> the capsule when the stem containing the coupler and about 1.5 cm <strong>of</strong> optical<br />

fiber is not supported on the shaker table.<br />

In a vibrationally noisy environment, the ratio <strong>of</strong> the acceleration sensitivity to the acoustic sensitivity<br />

to is a good figure-<strong>of</strong>-merit for comparison <strong>of</strong> their noise-limited performance. Since this ratio removes<br />

the output parameter (volts or radians), it is also useful for comparing fiber-optic hydrophones to<br />

piezoelectric hydrophones. Reporting this ratio in decibels, the hydrophone capsule discussed in this<br />

paper has a figure-<strong>of</strong>-merit is -134 dB re g/Pa. This is 43 dB better than the Naval Research<br />

Laboratory's planar flexible fiber-optic interferometric hydrophone and 57 dB better than the piezoelectric<br />

polymer (PVF2) hydrophone <strong>of</strong> similar geometry3. An overall comparison <strong>of</strong> the NRL and NPS designs<br />

is given in Table 2.<br />

TABLE 2. NPS and NRL Planar <strong>Hydrophone</strong>s<br />

Performance measurement NRL NPS NPS Advantage<br />

Normalized acoustic sensitivity (dB re: lpPa1) -321 -297<br />

Normalized acceleration sensitivity (dB re: lj.iPa1) -144 -163<br />

Acoustic-to-acceleration FOM (dB re: g/jiPa) -177 -134<br />

x<br />

Table 2. The NPS flexural disk hydrophone is 16 times more sensitive to pressure and 9 time less<br />

sensitive to acceleration than the NRL spiral3 giving the NPS design a factor <strong>of</strong> 43 dB improvement over<br />

the spiral in a vibrationally noisy application.<br />

.<br />

>(<br />

Freauenc, Kertz<br />

.<br />

:<br />

iiiIziiziizI Iz1 1ZI I 11 1<br />

cT .<br />

:<br />

1000<br />

(dB)<br />

+24<br />

+19<br />

+43<br />

26 1 SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990)


5. DISCUSSIONOF RESULTS<br />

The results reported here represent the highest normalized sensitivity ever reported for a fiber-optic<br />

hydrophone designed to operate under conditions <strong>of</strong> static pressure higher than 500 psi. It also represents<br />

the best acceleration cancellation figure <strong>of</strong> merit for any fiber-optic hydrophone. One additional feature <strong>of</strong><br />

the design which should be considered equally important is that its performance is in excellent agreement<br />

with simple elastic theory so that hydrophone designers can scale these results as necessary for other<br />

applications requiring differing materials, bandwidth, depth tolerance, etc., with a good deal <strong>of</strong><br />

confidence.<br />

The fact that this hydrophone capsule can be fabricated as a single unit cast out <strong>of</strong> epoxy suggests that<br />

its cost for mass production may be significantly lower that other hydrophone designs requiring machined,<br />

thin, air-backed, cylindrical mandrils which may require subsequent fabrication steps that make internal<br />

access for fiber winding, and thus push-pull performance impractical.<br />

Thus far, the discussion <strong>of</strong> acoustic performance has concentrated on the normalized sensitivity. That<br />

parameter provides a comparison between design strategies on a meter-for-meter basis. We would like to<br />

conclude this discussion with comparison <strong>of</strong> the NPS flexural disk and NRL flexible planar designs which<br />

also introduces the effects <strong>of</strong> "winding density" since this allows comparison <strong>of</strong> the actual available phase<br />

modulation (i.e., signal). The overall available phase modulation amplitude determines the detection<br />

threshold if one assumes a given demodulator detection threshold. Similarly, one could use an acoustic<br />

basis to set the detection threshold and allow the hydrophone sensitivity to dictate the required demodulator<br />

performance, which then determines the cost and complexity <strong>of</strong> the demodulator, laser diodes, etc.<br />

6. ACKNOWLEDGEMENTS<br />

Over the past six years this work has been supported by the Office <strong>of</strong> Naval Research - Physics<br />

Division, the Office <strong>of</strong> Naval Technology, the Space and Naval Warfare Systems Command, the Naval<br />

Sea Systems Command, and the Naval Postgraduate School Direct Funded Research Program.<br />

7. REFERENCES<br />

a) Lieutenant, U. S. Navy<br />

1 . T. G. Giallorenzi, A. Dandridge, and A. D. Kersey, "Interferometric and intensity sensors become<br />

more sophisticated" Laser Focus World 27(7), 137-143 (July, 1989).<br />

2. N. Lagakos, A. Dandridge, and J. A. Bucaro, "Optimizing fiber coatings for interferometric acoustic<br />

sensing (U)", U. S. Navy J. Underwater Acoustics 37(3), 233-254 (July,1987).<br />

3 . N. Lagakos, P. Ehrenfeuchter, T. R. Hickman, A. Tveten, and J. A. Bucaro, "Planar flexible fiberoptic<br />

interferometric acoustic sensor", Opt. Lett. 13(9), 788-790; OFC '90 Tech. Digest (22-26 Jan<br />

1990, San Francisco, CA) p. 46; N. Lagokos, J. A. Bucaro, P. A. Ehrenfeuchter, T. R. Hickman,<br />

A. B. Tveten, and A. Dandridge, "Planar-conformal fiber optic acoustic sensing element", in <strong>Fiber</strong><br />

<strong>Optic</strong> Smart Structures and Skins II, Proc. Soc. Photo-<strong>Optic</strong>al Instrumentation Eng. (SPIE) 1170,<br />

472-477, (1989).<br />

SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990) / 27


4. G. B. Mills, S. L. Garrett, and E. F. Carome, "<strong>Fiber</strong> optic gradient hydrophone", in <strong>Fiber</strong> <strong>Optic</strong> and<br />

Laser Sensors II, Proc. Soc. Photo-<strong>Optic</strong>al Instrumentation Eng. (SPIE) 478, 98-103 (1984).<br />

5 . G. B. Mills, "<strong>Fiber</strong> optic gradient hydrophone", Master's thesis, Naval Postgraduate School (June,<br />

1984), DTIC Report No. A 143 975.<br />

6. G. E. MacDonald, "<strong>Fiber</strong> optic gradient hydrophone construction and calibration for sea trials",<br />

Master's thesis, Naval Postgraduate School (March, 1985), DTIC Report No. A 156 469.<br />

7 . P. A. Feldmann, "Construction <strong>of</strong> a fiber optic gradient hydrophone using a Michelson<br />

configuration", Master's thesis, Naval Postgraduate School (March, 1986), DTIC Report No. A 167<br />

892.<br />

8. D. L. Gardner, T. H<strong>of</strong>ler, S. R. Baker, R. K. Yarber, and S. L. Garrett, "A fiber-optic<br />

interferometric seismometer", J. Lightwave Tech. 5(7), 953-960 (1987).<br />

9. D. L. Gardner, and S. L. Garrett, "<strong>Fiber</strong> optic seismic sensor", in <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors V,<br />

Proc.Soc. Photo-<strong>Optic</strong>al Instrumentation Eng. (SPIE) 838, 27 1 -278 (1987).<br />

10. D. L. Gardner, "A fiber-optic interferometric seismic sensor with hydrophone applications", Ph.D.<br />

thesis, Naval Postgraduate School (September, 1987), DTIC Report No. B 1 12 487<br />

1 1 . D. L. Gardner and S. L. Garrett, "High sensitivity fiber-optic compact directional hydrophone (U)",<br />

U. S. Navy J. Underwater Acoust. 38(1), 1-23 (1988).<br />

12. D. A. Brown, T. H<strong>of</strong>ler, and S. L. Garrett, "<strong>Fiber</strong> optic flexural disk microphone", in <strong>Fiber</strong> <strong>Optic</strong><br />

and Laser Sensors IV, Proc. Soc. Photo-<strong>Optic</strong>al Instrumentation Eng. (SPIE) 985, 172-182 (1988).<br />

1 3 . D. A. Brown, "A fiber-optic flexural disk hydrophone", Master's thesis, Naval Postgraduate School<br />

(March, 1989), DTIC Report No. B 132 243.<br />

14. J. A. Flayharty and B. M. Fitzgerald, "Design and fabrication <strong>of</strong> a fiber optic acceleration canceling<br />

hydrophone", Master's thesis, Naval Postgraduate School (September, 1989), DTIC Report No. B<br />

136 179.<br />

1 5. D. A. Danielson and S. L.Garrett, "<strong>Fiber</strong>-optic ellipsoidal flextensional hydrophones", J. Lightwave<br />

Tech. 7(12), 1995-2002 (1989).<br />

16. D. A. Brown, S. L. Garrett, and D. A. Danielson, "<strong>Fiber</strong>-optic oblate flextensional hydrophone", in<br />

<strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VII, Proc. Soc. Photo-<strong>Optic</strong>al Instrumentation Eng. (SPIE) 1169,<br />

240-248 (1989).<br />

17. D. A. Brown, T. H<strong>of</strong>ler, and S. L. Garrett, "High sensitivity, fiber-optic, flexural disk hydrophone<br />

with reduced acceleration response", <strong>Fiber</strong> and Integrated <strong>Optic</strong>s 8(3), 169-191 (1989).<br />

1 8. K. Wetterskog, B. L. Beaton, and J. Serocki, "A fiber-optic acceleration canceling hydrophone made<br />

<strong>of</strong> a castable epoxy", Master's thesis, Naval Postgraduate School (June, 1990).<br />

19. M. R. Layton, B. A. Danver, J. D. Last<strong>of</strong>ka, D. P. Bevan, and E. F. Carome, "A practical fiber<br />

optic accelerometer", in <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors V, Proc. Soc. Photo-<strong>Optic</strong>al Instrumentation<br />

Eng. (SPIE) 838, 279-284 (1987).<br />

20. M. R. Layton, A. D. Meyer, D. P. Bevan, and J. S. Bunn, "Multiplexed fiber optic hydrophones",<br />

DOD <strong>Fiber</strong> <strong>Optic</strong>s Conference '88 (22-25 Mar 1988, McLean, VA) p. 85.<br />

21. T. J. H<strong>of</strong>ler and S. L. Garrett, "Thermal noise in a fiber optic sensor", J. Acoust. Soc. Am. 84(2),<br />

471-475 (1988); J. Acoust. Soc. Am. 87(3), 1362-1365 (1990).<br />

22. T. Sullivan and J. Powers, "Piezoelectric polymer flexural disk hydrophone", J. Acoust. Soc. Am.<br />

63(5), 1396-1401 (1978).<br />

28 / SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990)


23. D. Ricketts, "Model for a piezoelectric polymer flexural plate hydrophone", J. Acoust. Soc. Am.<br />

70(4), 929-935 (1981).<br />

24. S. Timoshenko and S. Woinowsky-Kreiger, Theory <strong>of</strong> Plates and Shells, 2nd edition (McGraw-<br />

Hill, 1959), pp. 51-58.<br />

25. C. L. LeBlanc, Handbook <strong>of</strong> <strong>Hydrophone</strong> Element Design Technology, (Naval Underwater<br />

Systems Center Technical Document 5813, 1 1 October 1978),<br />

26. A. E. H. Love, A Treatise on the Mathematical Theory <strong>of</strong> Elasticity, 4th edition (Dover, 1944), p.<br />

484.<br />

27. United Resin Corporation, 2730 S. Harbour Blvd., Santa Ana, CA 92704.<br />

28. S. L. Garrett, D. A. Brown, J. Serocki, K. Wetterskog, and B. L. Beaton, "Elastic moduli <strong>of</strong><br />

castable epoxies", submitted to the U. S. Navy Journal <strong>of</strong> Underwater Acoustics.<br />

29. S. L. Garrett, "Resonant acoustic determination <strong>of</strong> elastic moduli", J. Acoust. Soc. Am. 88(1), 210-<br />

221 (1990).<br />

30. Mechanical Measurement Devices, Model 3012A Dynamic Modulus Measurement System, P. 0. Box<br />

8716b, Monterey, CA 93943.<br />

3 1 . ASTM D 638, 1989 Annual Book <strong>of</strong>ASTM Standards,(American Society for Testing and Materials,<br />

1989), Section 8, Plastics, volume 08.01, pp. 156-166.<br />

32. Coming Glass Works, Telecommunications Prothicts Division, Coming, NY 14831.<br />

33. A. W. Leissa, Vibration <strong>of</strong>Plates, NASA Tech Doc: SP-160-N70-18461 (1969),<br />

34. Interfuse 945 Series single mode coupler, Amphenol Corp., Lisle, IL 60532.<br />

3 5 . Seastar <strong>Optic</strong>s, Model PT-450, Sidney, B .C., Canada V8L 31.<br />

36. Litton Guidance and Control Systems, Woodland Hills, CA 91367.<br />

37. Hexcel Corporation, Chemical Products Division, Chatsworth, CA 91311.<br />

38. International Transducer Corporation, Goleta, CA 93117.<br />

39. R. J. Bobber, Underwater Acoustic Measurements, (Naval Research Laboratory, 1970).<br />

40. J. Lighthill, Waves in Fluids, (Cambridge University Press, 1978),<br />

41. J. D. Ramboz, "Calibration <strong>of</strong> pickups", in Shock and Vibration Handbook, 2nd ed., C. M. Harris<br />

and C. E. Crede, editors (McGraw-Hill, 1976), pg. 18-14.<br />

42. M. B. Johnson, "Reciprocity calibration in a compliant cylindrical tube", Master's Thesis, Naval<br />

Postgraduate School, Monterey, CA, June, 1985; DTIC Report No. A158-998.<br />

43. D. A. Brown and S. L. Garrett, "An interferometric fiber optic accelerometer", in <strong>Fiber</strong> <strong>Optic</strong> and<br />

Laser Sensors VIII, Proc. Soc. Photo-<strong>Optic</strong>al Instrumentation Eng. (SPIE) 1367-34, (1990), in<br />

press.<br />

SPIE Vol. 1367 <strong>Fiber</strong> <strong>Optic</strong> and Laser Sensors VIII (1990) / 29

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!