LJBflARY TECHNICAL REPORT SCCTKXI NAVAL POSTGRADUATE SCHOOL MONTEREY. GAUFOOMIA S&940 NPS-61Md73111A NAVAL POSTGRADUATE SCHOOL // Monterey, California 15 Nov 1973 NPS-61Md73111A Pred icting Sound ] Dhai se and Amplitude Fluctuations due to Micros tructure in the s Upper Ocean Physics & Chemistry H. Medwin Department Approved for public release; distribution unlimited. FEDDOCS D 208.14/2:NPS-61MD73111A NAVAL POSTGRADUATE SCHOOL Monterey , California Rear Admiral Mc B. Freeman M0 U. Clauser Superintendent Provost TITLE: Predicting Sound Phase and Amplitude Fluctuations Due to Micro- structure in the Upper Ocean AUTHOR: H. Medwin ABSTRACT: Tne temporal and spatial variations of the index of refraction cause fluctuations of sound phase and amplitude that can be completely understood only by defining the index in terms of the duration, location, range and time of the acoustic experiment. A truncated "universal" spatial correlation function of the index has been derived from a simplified form of the Kolmogorov-Batchelor spectrum of temperature fluctuations in a homogeneous, isotropic medium. Although this correlation function is shown to be predictable simply from the depth of the experiment, it is of only limited validity with respect to large spatial lags. However, a Gaussian extrapolation of the "universal" correlation function together with the standard deviation of the index provides simple useful predic- tions of the sound fluctuations due to temperature micros tructure in the upper ocean. This task was supported by Naval Ship Systems Command (Code PMS 302). TABLE OF CONTENTS 1. Introduction: Sources of Sound Phase and Amplitude Fluctuations 1.1 Microstructure in Mixed Water 1.2 Layered Microstructure 1.3 Other Underwater Structural Features 2. Wave Number Spectrum and Spatial Correlation of the Index of Refraction in a Mixed Layer 2 . 1 The Wave Number Spectrum 2.2 The Spatial Correlation and Structure Functions of the Microstructure 3. Sound Phase and Amplitude Fluctuations Due to Microstructure 3 . 1 Phase Fluctuations 3.2 Amplitude Fluctuation 3.3 Small Amplitude Fluctuations: Laboratory Experiments 3.4 Large Amplitude Fluctuations 3.5 Fluctuations at Sea 3.6 The Limits of Predictions 4. Conclusion References Initial Distribution List Form DD 1473 1 . Introduction Sources of Sound Phase and Amplitude Fluctuations As it proceeds away from its source, a sound at sea encounters changing values of temperature, salinity and density, as well as entrained objects and bubbles in motion. These acoustically significant quantities are constant neither in space nor in time. Slower than the microscale fluctuations due to surface waves and turbulence of the medium, which show perceptible changes during a time scale of the order of seconds , are changes due to internal waves with periodicities of the order of minutes , tidal and diurnal variations , and seasonal changes. Depending on the duration of the study, some of these fluctuations may appear to be simply periodic, some show the spectral char- acteristics of a narrow band noise, many are describable only by statistical methods , and still others, such as the appearance of schools of fish or the incidence of storms, are largely intermittent in character. (Weston 1969). It is important to make the point that it is not the inhomogeneities but rather it is the temporal change in the position or character of the inhomogeneities that is the source of interest in this report. The phase of a sound wave travelling through a stationary medium depends on the speed of sound along the path. This dependence is expressed by equations 1.1 and 1.2: ^^ ds = «• ^fiw ds (x) has been shown to be (Batchelor 195 6) uniquely defined by the two parameters, the kinematic shear viscosity, v , and the rate of supply (or removal) of energy, e . The equation for the one dimensional PSD is *v(x) = bx"5/3 (2.1) where b = b {v , e) x = 2tt/A= wave number of a turbulent velocity component A = effective wave length of a turbulent velocity component € = energy dissipation, per unit mass, per unit time V = kinematic shear viscosity of water The -5/3 rds power law has been amply verified for components of turbulent velocity in laboratory experiments and, to a lesser extent, at sea. We are interested in the turbulent velocity not only because it is a significant source of sound fluctuations in special situations such as turbu- lent channels but, more importantly, because it is a guide to the behavior of temperature variations. And temperature variations are generally the principal sources of sound fluctuations in the ocean. The temperature often acts as a convected, passive contaminent which closely follows the turbulent velocity behavior. The spectrum of temperature fluctuations has been studied for such a case (Batchelor 1959) and again the -5/3 law is predicted for the inertial subrange. The predictions have been largely verified (Stewart 1962, Grant 1968) by ocean measurements. 4 o Fig. 2 Temperature and velocity spectra at depth 2 7m in the open sea, from Grant (1968a). The predicted transition wave number, x. /is shown, as calculated from Eq. (2.8). Fig. 2 plots the spectra (Grant 1968) of a component of velocity and tempera- ture measured at 27m depth, compared with the theories described in the references. The figure shows the extension of the Kolmogorov theory from the inertial subrange, where the -5/3 slope is confirmed, into the higher wave number dissipation sub-range. The boundary between the two subranges is often assumed to be at the Kolmogorov wave number, = H b = (e h3)k (2.2) Values of € , derived from measurements of the turbulence spectrum, are a function of depth and have been found (Grant 1968) to be as high as 0.5 cm sec atjan acean depth of 15 meters in well-mixed water and as small as 1.7 x 10 cm in quiet water below the mixed region at 213 meter depth. Looking again at Fig. 2 we see a rise in the temperature spectrum, also predicted Jpy BatcheLpr (1959). At this depth, 27m, £ was determined to be 5.2 x 10 cm /sec . Assuming v = 1.4 x 10 " cm /sec, Eq. 2.2 yeilds x =6.6 cm . The rise in the temperature spectrum occurs between x and S _2 ° x —2.4 x 10 x . At greater depths, the dissipation constant and the Kolmogorov wave number are generally smaller. Having identified the high wave number end of the isotropic -5/3 spectrum, we now seek to establish the low wave number limit. Unfortunately, at the low wave number end of the spectrum of temperature micro structure, the PSD is very dependent on the local conditions and history of the water mass, and only empirical results are available to guide us. A useful, but very crude, generalization that can provide some guidance is that the largest temperature wave length that could conceivably be isotropic at the depth H would be of extent 4H. The corresponding isotropic wave number would be 2TT TT in i\ This would suggest that for Fig. 2, where H = 27m, the lower limit of the -5/3 spectrum might be at Hm = 5*8 x 10~ cm_1 (°r lo8io Hra =-3-2). In fact, the data of Fig. 2 show that at values of x much smaller than 5.8 x 10~ (perhaps at x- 0.02 cm" ) both the temperature and the velocity PSD break away from the -5/3 slope. This suggests that, realistically, the lower 10 end of the -5/3 spectrum to be expected at sea is somewhere between x and x . We will call that transition wave number, x . o t The appropriate variable to describe sound phase and amplitude fluctuations will turn out to be the speed of sound, and so it is the spectrum of fluctuations of the speed that we must seek. The connection is direct. For simplicity we will assume that the speed depends only on the temperature. We have previously seen that the fluctuations in temperature micros tructure are much less than the average temperature. The fluctuations in the speed of sound are also a very small fraction of the average value. It therefore turns out to be most useful to define not only the index of refraction of the sound speed, at position U — ♦ ft(l)s _i^_ (2 < 4) — + but also the excess index of refraction, \i (R) , which will be the particular parameter for all future discussions of fluctuations. H (R), c <»>" ic <*)> = n ft. k< ! The mean value, ( ) , is to be calculated over the time of the experiment. It is essential to realize that [± (R) is sensitive to position R* . However, for typographic convenience we will drop the notation for the explicit dependence on position so that henceforth we write p. (R) =|i. Further, we define / \ a * - I 2\z (2-6) (p,) = o and a = \\t, ) The point must also be made that if the time or duration of our averaging process is changed, the value of <|_i) and a will change. We must immediately accept that these quantities must have their temporal limits restricted. A logical criterion for the restriction is that for the present we are concerned only with locally homogeneous regions of the ocean, only for times comparable to the duration of the acoustic experiment. We will be more precise about these limitations , shortly. Since the excess index, \i, is proportional to the excess speed of sound, which in turn is assumed to be proportional to the excess temperature, AT, a spectrum of M- can be expected to resemble the spectrum of AT. It will have an inertial subrange in which the log of the power spectral density, $ (x) will have a slope of -5/3 when plotted against log x , as in Fig. 2. That isotropic inertial subrange will have an upper limit x* and a lower limit which we will call x^. 11 The lower bound of the isotropic inertial subrange of temperature x appears to be approximately a function of depth, H, alone if we define t it as \ = °-5 (*o Km> - 0.3 (V/3 < TT 2H (2.7) Fig. 3 is a graph of x^ calculated from Grant's (1968) values of c and H, as well as those of Nasmyth (1970). O GRANT etal (1968) X NASMYTH (1970) 20 40 60 80 DEPTH, H METERS 00 Fig. 3 Transition wave numbers, calculated from Eq. (2.7) using data of Grant (1968a) and Nasmyth (1970). The empirical fit to the data, given by Eq. (2.8) is shown. .The equation of the best simple approximation to this graph, is of the form -H/h x = B e (2.8) where B = 9 .0 m-1 h = 40.0m H = depth, m. x = lower wave number limit of isotropic temperature spectrum, m 12 For example, for the experiment at 27M depth, shown in Fig. 2, Eq. (2.8) allows us to diagnose that the medium was no longer isotropic for x (x) >bx ' for x* ^x^x , and a rapid exponential drop-off for x>x . We will generalize and use this spectrum in the next section. The data available show that x «x. «x . o t m 2 .2 The Spatial Correlation and Structure Function The alternative to the spectral description of fluctuations of temperature or index of refraction is either the correlation function or the structure function, 13 which is simply derived from the spatial correlation. These functions turn out to be the most direct way to calculate the sound fluctuations , as we will see in Section 3 . One of the earliest determinations (Liebermann 1951) of the spatial correlation of temperature fluctuations was conducted by mounting a therm- istor on a submarine operating at a depth of approximately 50m. The standard deviation of the temperature fluctuations was found to be approxi- mately 0.4C at this depth (time of day, weather conditions unreported). The spatial correlation could be fitted approximately by the exponential law (shown by solid line in Fig. 5) cT <§) 2 = e -S/a (2.9) where a = 60. cm which thereby defined a correlation length, sometimes called a "patch size of 60 cm. i.o k^ \:i Kolmogoroy N^_ 0 ■*••• 100 120 E cm Fig. 5 Temperature correlation function from Liebermann (1951) with data (circles), empirical fits to exponential (solid line), and Gaussian (dashed line, correlation functions. The correlation based on Fig. 4 with x = 0.24 cm is shown by dotted line, identified as "Kolmogorov1 Others have used the Gaussian correlation function 2 CT (4) =e -(£/*)' (2.10) to describe micros tructure fluctuation. A Gaussian correlation function (dashed line) has been drawn in Fig. 5 for comparison with the exponential functions and the data (circles). 14 Both functions define the same correlation length, since they both reach e at £= a. The correlation function of the index of refraction will equal that of the temperature, if other parameters affecting the speed of sound have negligible effect, and if the fluctuations are small. Then cT(*)=cu («> It is not always possible to determine the correlation function C (£) or C (4) by measuring it directly. In any event it would be a boon if we could use the known universal characteristics of ocean micros true ture to deduce the spatial correlation function in the real world outside of the laboratory. Although this turns out to be not completely feasible, there are certain generalizations about the spatial correlation function that can be deduced; one of these is its form at small values of £. The spatial correlation function of the excess index of refraction can be obtained from the wave number spectrum by the Fourier transform theorem. In three dimensions this relation is s(?)-"-f- ^ y*)ei*'* dH (2-n) where f* is the vector spatial lag. If we now assume that the changes of the index of refraction are independent of the direction being studied we can simplify the integration. In spherical coordinates H»r. = urease- and dn = 2tth sin©- dn d & Then C (r) = -*Z— fQ K2 dH j" s (h) sin@r ^nr*** rf a ^ and the £ integration yields We now define the one dimensional spectral function, $ (x) for the isotropic medium "• 2 S^ (h) dn = 4tt h S^ (h) dH = § (h) dn and change (2.12) notation, r -» 4 # for consistency with previous work, so that 15 % (§) = —\- £ li£ri V (K) dH- <2-13> ■ £ may be in any assigned direction for the isotropic medium. In order to integrate (2.13) we start with the spectral conclusions of the previous section (Fig. 4). We are prepared to make four simplifying assump- tions in order to proceed: 1) $ (H) = $ + s(H - K ) for K < K , : * M- m my ra (2.14) $m;-$o Source Subrange where s is the slope constant= xm $m = peak isotropic PSD; $0 = $ (°) 2) * 00 = constant = 4 for * < * < ** Transition Subrange K -5/3 ra x 3) §^ (h) = $m ( t/H) for Ht < k < Kq Inertial Subrange 4) $ 00 = o for H > H0 Dissipation Subrange The approximate spectrum that follows these assumptions is shown by the dashed line of Fig. 4. The effect of the simplifying assumptions is to replace the highly variable anisotropic range and the fixed, but complicated, isotropic range by a fixed, easily-integrable , spectrum. This will permit us to assess the relative impor- tance of each of the sub-ranges and to determine what universal generalization can be made, (if any) about C (£) at sea. Using the assumed spectrum of Eq. (2.14) it is possible to immediately determine the excess local index of refraction, a in terms of the peak spectral density, $ , and the constants x and x . Recalling that the mean squared value of any fluctuating quantity may be obtained by integrating the spectral density over all values of the appropriate parameter (frequency or wave number) we find (V00 (u ) = J $ (K) dK ^r O JJ, m m t Since x « x and & < & , the result of the integration simplifies to m t o m (j, = 5/2 k $ t m 16 It is interesting to determine the parts of the spectrum most responsible for this variance of the fluctuations . The terms in the integration were 1/2 x <£ source subrange m m x, $ transition subrange t m 3/2 xx $ inertial subrange t m Since x « x we conclude that the contributions to cr come almost equally from the transition and the inertial subranges. Furthermore, our simplifying assumption of the particular form of the source subrange can have little effect on (j because that contribution is proportional to x , and x «x . \i m m t A second general conclusion can be reached if we restrict our interest to very small values of the spatial lag, £ . Then we can substitute the expansion for the sin x£ and integrate Eq. (2.13) with ease — Jo $ (h) dx a__ (J H [$ + m (h - x )] dx + Jk A m m § dx nv m + jHoX2J (^)5/3dK K m x 2 The first integration yields (\i >, as we saw above, so that the first term is unity. The integration over the three terms in the curly bracket is also simple so that .2 . 5/3 4/3 C (§) = 1 - % $m Kt ^— r2 (2.16) 8 o- and with the aid of (2.15) C (§) = 1 - 0.05 k 2/3 x 4/3 §2 for x 6«1. (2.17) fJL tO O The significance of this result appears if we consider any specific depth ^ say 50m. Then (2.3) gives x = tt/10 cm" , (2.8) predicts x = 2.4 x 10~ cm" and therefore, at 50m depth, we can expect C (§) = 1 - O.J468 §2 17 For small displacements this is an expansion of the Gaussian form C^ ' V- o SlvEH § *« + S (H - H )]dH + J ^V^»'dH + J °sAlLHi* ^5/3 , M- U H = o H§ u in v m/J h h£ m k hZ m x m t The coordinate for integration is now changed to 0 = x £ to « ■ i/:l -&**\ ♦ -f - *» * h:! 5<^>v ♦ ?2i:^*/H m 3 ' s in R R £i The approximations are now made for J p d/S = |9- -~r-, + 5°5 , in the source integral and in the transition integral. Because of the rapid convergence due to the jg 8/3 in the denominator, the third integral is comfortably approximated by using the sine expansion from zero to tt/2 and again from ir/2 to n. Using (2.15) the result of these integrations, valid for £ £— with about 5% error, is \ C (i) = 1.0 - 0.48(xt£)2/3 +0.28 (x^)2 +6.7 x 10~4(xt£)4 (2>18) The curve is shown in Fig. 6 I.O- *-* \j 0.8- v: 3. a6- ^ 0.4- 0.2- . 0 i i i i i i w 0 05 1.0 1.5 2.0 2.5 3.0 M Fig. 6 Spatial correlation function C (x £) derived from spectrum of Fig 4 It turns out that the source subrange makes essentially no contribution to the correlation function for 4 = — J dx = — J dx c o c o o o because the average index, (n) , is unity. 20 It is the deviation from the mean that we must calculate: At = t - =r — J n (x,T) dx — J dx CO CO o o i rR = — J0 M.(x,T)dx where p, (x,T) = n (x,T) -1 From this we get the square of the deviation from the mean (At) =^ — „) ( Jo |jdx f^dx) which can be written (At) =— J^ J^ |j (x^T)^ °(X2'T) o dxldx2 2 1 R R The mean square deviation from the mean is (At > = - dxn f " (u (xn ,T) /m\\_i cz«Jol«Jol [i (x2,T)> dx2 o The integrand can be written in terms of the spatial correlation of the excess index of refraction C (£): = (n*> C^ (x2-Xl) = a* C (VXl) . a2 „ /xR 2 R I In these terms (At ) = — ^ J C (x^-xjdx, 0 o ^ / 1 2 v p *. 4. (3.3) o We seek the mean squared phase fluctuation which is calculated by substituting cp = ceAt where (pis the instantaneous deviation of the phase from the value due to the mean speed, c . The units of

=a2k\fc (4) d£ D«l (3.4) 2 2 2 2 Since (cp } = (cp> + Var cpand (cp> = o, by definition, (cp )= Var (p= a . (3.5) We have only to know the spatial correlation of the index of refraction, from the source to the receiver, in order to carry out the integration, and obtain the variance of cp . 2 2 It should be pointed out that the quantity a = ([± > must be known at the time, and location of experiment. The significance of this comment is that, in the ocean, the evaluation of a can be strongly dependent on position and the range at a given time, or a function of the time and duration of measurement. 21 For example during passage of an internal wave, at positions in the thermo- cline there will be large variations of the temperature and the speed of sound, over times of the order of several minutes. It is also convenient, sometimes, to assume that the spatial correlation function drops to zero after a few correlation lengths in either direction from any reference point. Then for the sound path of length R »a we can extend the integration to infinity in both directions , or double the integration from zero to infinity, without affecting the total integration, 00 a2= ?1 as a requirement for far field propagation, the fluctuations of the medium, a must be so small that 2 2 k* a. aR « x In practice this condition is not difficult to fulfill; field measurements generally show q <10 22 The Gaussian correlation function C (£) <=e "^/a) (3.8) is attractive because of the simplicity in evaluating the integrals of Eqs . (3.6) and (3.7). For the Gaussian correlation function the phase fluctuations are: a2 = = ^CT2 k2 Ra D«l (3.9) 2 . 2V y? 2 2 (3.10) and a = < =-r— a k R a D» 1

= a2 K2 Rfc (£)d£ D»l (3.14) A (_i Jo (j ^ ^ where cr is the fractional standard deviation of the pressure amplitude o = <(P-PQ >2)% (3.14a) A Po 23 Comparison of expressions for amplitude fluctuations with those for phase fluctuations is revealing. For example, for the high frequency situation D «1 the mean squared amplitude fluctuations depend on an integration of the fourth derivative of the correlation function in the transverse directions, y and z, rather than a straightforward integration over the correlation function in the x direction as in Eq . (3.8) for the phase shift. This means that the amplitude fluctuation is caused by the curvature (really double curvature) of the correla- tion of the refractive index, transverse to the sound path. The dimpled medium in the case of amplitude fluctuations therefore behaves as a series of lenses coverging or diverging acoustic energy along the sound path (notice the evalua- tion at zero displacement in the y and z direction, that is, along the beam). At the other extreme of propagation parameters, however, when the range is larger, or the frequency is lower, such that R/a » ka (D »1), the sound field fluctuations are caused by field distortions due to a large number of inhomogeneities along the path in the medium; these distortions interfere, constructively and destructively along the entire path and this interference effect is common to both phase and amplitude. For this condition the mean squared fluctuations of phase and amplitude are identical. Both are linearly dependent on: a) the distance of propagation, b) the frequency squared, and c) the integrated spatial correlation function. It must be pointed out that all of the derivations to this point have been based on the assumption of single scattering by weak fluctuations. The consequent solutions are valid only in so far that very long paths and very large fluctuations are not allowed. In practical propagation over very long paths the magnitude of the phase and amplitude fluctuations cannot increase with R without limit, or they would reach the absurd conditon of fluctuations that are greater than the quantity itself. In fact, "saturation" must take place and the increase of the fluctuation with range must reach a limit that will be indicated, shortly. If the Gaussian correlation function of the index is assumed (Krasilnikov 195 6) the variance, ah, is: °l - C-^fV) k' R a <1-^-tan"1 D)1 (3.15) al (X2> = (-^T") °l (R/a>3 D = (~T~) °l ^ (Ra) D >;> l (3'17) 24 3. 3 Small Amplitude Fluctuations: Laboratory Experiments 2 The dependence of small values of or on the parameters of the medium and the frequency has been put on a firm experimental basis in a beautiful set of laboratory studies (Stone and Mintzer 1962) in which the micros tructure of the index of refraction was obtained by heating the tank of water through which the sound beam was propagated. First the rms value and the spatial correlation function of the temperature microstructure were obtained by direct measurement. From these data the value a / and the correlation function, Cyi (£ ) , of the excess index of refraction were calculated. The spatial correlation function was fitted to the Gaussian form, and the correlation length, a, was calculated from Eq. (3.8). The measured temperature micros tructure , fitted to the Gaussian correlation function showed a correlation length, a = 3.5 cm. Additional comparison showed that the rms excess index of refraction, deduced acoustically, was approximately 25% less than the value 1.6 x 10 , calculated directly from temperature measurements. One flaw was revealed in the experiment: the acoustically determined value of the assumed Gaussian correlation length was only about one-third of the value (3.5 cm) directly measured by thermistors. This discrepancy was possibly due to the error of assuming that the correlation function was Gaussian. As shown in section 2, the Gaussian form is appropriate only for very small displacements. A combination of Gaussian and Kolmogorov correlation functions would probably have been a better description of the medium for insertion into Eqs. (3.13) or (3.14). 3. 4 Large Amplitude Fluctuations - Large Ranges So far we have considered small fluctuations for two cases: 1) Change of sound pressure amplitude, P, in the presence of zero (or negligible) phase shift. In section 3.2 we gave expressions for 2 P -P 2 2 cr. = // o_. v . The type of pressure field at x = R that results in = 0; S o b (x,t) = time varying magnitude of the scattered pressure which is 90 out-of- phase with the original wave at x. We assume b is Gaussian distributed and = 0. b *A. s s Equation (3.18) resolves the fluctuation scattered pressure, P of Fig . 7 into components in-phase with the depleted original wave A (x) ana out-of- phase with it. A typical situation in Fig. 7 shows that «2 2 /I P = a + b s s s To get the typical amplitude, Pf of the total wave at x as a function of A, a , b , Eq. (3.18) is rearranged to P = (A + a ) cos (cot - kx) + b sin (wt - kx) (3.19) = P cos (o>t - kx - $) where P = pressure amplitude at x (3.20) r/» x2 u 2-,l/2 = [(A + a ) +b ] s s <$ = phase angle with respect to original -1 bs wave at x = tan ( ) (3.21) A+a s It is useful to expand _P and $ for small fluctuations a — b «A. Using the binomial theorem for _P -A(l +-5"- +%A1N )-A+as (3-22) and the small angle approximation for $ I = tan"1 (-^- ) - A+a ; A (3.23) s 27 2 It is therefore clear that for small fluctuations the = < P2) UJ > = <(A4.s)2 + b/) ■= + (2ao A) + (a 2> + _ 9 9 9 ^ ^ Ao " * + (as > + (3.28) To determine the attenuation rate, a , we use (3.2 7) and (3.28) and form S3 An2 -A2 _ (as2) + (bs2> _ , -20i R A 2 " A 2 " L "e S (3.29) A A o o The evaluation can be made in the small fluctuation range where the ingredients of (3.29) are well known. There, using the expansion of the exponential for small argument and (3.24) and (3.2 5) aA + a9 = 2asR (3.30) The condition for long range is of particular interest. Then we have, from (3.7) and (3.14) °l = al >D>>1 <3-31> so that the attenuation constant of the original beam due to energy going into fluctuations , is * =a2/R ,D» 1 (3.32) In particular, when the scatter is due to a medium with a Gaussian correlation function, C (£), we have, using (3.17), a =-^-a2K2a (3.33) s 2 |i 29 We can now proceed to consider large ranges and large fluctuations. The quantity observed in any experiment is the pressure amplitude P. Eq. (3.2 0) shows P to be a function of all three quantities: The steady coherent component, A, and the fluctuating random components, a and b . The conventional way to express the amplitude variations is in the ratio (CAV)2= -2

2 (3-34)

2 where CAV = coefficient of amplitude fluctuations. It might be noted that for small fluctuations we have simply CAV = a., Eq. (3.14a). Here, for large fluctuations we find, from the conservation of energy assumption that = = Ao2 (3.35) 2 But the calculation of

involves a more involved evaluation of the mean value of P given by Eq. (3.2 0) when both a and b vary randomly, each with an assumed Gaussian probability density function. The mean value of the sum has been found by Brownlee (1973) and is

=^AQ [1 -e-2asR}HexP[-2/(e2^R-i)]}^ £ ^ i •>, , (3.5)b2 (3.5 7)b3 (3.36) where £ = 1 + Jb + ^, .j. + tjj v* + b « fcCe2** -D"1 2 2 When this is now squared and combined with P in Eq. (3.34) the A cancels and the result is expressible solely in terms of the attenuation constant, a , and the length of path, R. (CAV)2 =-1 +-^exp ( ,l„ m . ) {[1 -exp (-2« RJIS2]}"1 (3.37) 77 exp(2asR)-i s This messy expression has a simple maximum value for large fluctuation and large ranges; we get (CAV)2 = -1 + -4" = 0.273 for 2a R » 1 v tt s (3.38) or (CAV) = 0.522 max The complete behavior of the coefficient of amplitude variation is shown in Fig. 8. 30 VM i -1-1 4-J d o fO 3 4-> i—4 o 3 CD 1 — 1 T3 >4-l 00 3 4-> -•-1 a) d a H £ (0 C o O IS c -C o 0 •r-4 CO 4-> OJ TJ S-4 -1-1 (0 > N T3 ffi (0 -a ^ d LO -a CM c 4-> m rd 4-> CO r^ 0) CD to jC cr> 6S 4-> 4-4 r— 1 O O c ,* 0 £ d -1— 1 4-> o d 4-1 of Shvac ure only IO co C data truct d o -r-4 4-> CO cD o x: u, H O 1-4 -i-H rd . S > ^o " CD co 3 -I-H c -o i— 1 o a, m-j cD 6 « -Q rd 3 0 ^■M <-M o *- d O 4-> — i CD C 0) •^H o fO co ■1-1 C CO 4-4 o :> a) 0 O GO • Cn PL, o b 31 which is the graph of the square root of (3.37). We observe that the fluctuations (CAV) = a for small values of a R, that is, for small fluctuations. The linear dependence of (CAV) on range continues until approximately a = 0.35 after which growth slows. (CAV) reaches 90% of its ultimate value at approximately VK (J2 M can be determined for any spatial lag, the value of C (£ ) at £ >R is quite irrelevent to a calculation of the sound fluctuations for a path of extend, R. Therefore practical acoustical application of integrals of C (£)d£ will be satisfied by an upper integration limit, R, rather than °° . The microstructure study of section 2.2 leads us to have confidence in the form of C (£) only for spatial lags over which the Kolmogorov spectrum is dominant. We have set these limits as 0 <£ o o o x: jo U CD ^_, CO o (D x: t*» w o g 6 c a p CD CD w (0 en CD fd fd CO 0 S-. p -M u p ii CO O S-. O XS CD +-> fO o c p +J CD Xi +j CO •iH CD p T3 CD o o cp r-> CD O T3 CO 6 H CD > o . (D 0> g Q_ lO ^ en C-> _e 2 * c CD £ CD 3 en fO CD c 6 .2 - -t-> fO CD CD o (D a en > O cn o .2 £ -3 r— I (0 O ^ -S o o o CD ^-. a co X j3 (D CO Z! CD C (0 -rH en en P fO ,Q O •■H 33 it levels off while the actual correlation function shows a continuing decrease and then an oscillation with small amplitude around the zero correlation axis. In order to obtain the integral of C (£) for a propagation experiment in such a sea we have two choices: One solution would be to determine C (£) and C (£) out to the range of the acoustic experiment; this would not only be a formidable task but unrealistic for any long range experiment. The second, more expedient, choice is to accept a judicious combination of theory and empiricism. We propose to use the theoretically based microstructure of Fig. 6 to determine the correlation length, a, and then to use the Gaussian correlation function to approximate the true correlation curve. Judging by Fig. 9, this will result in slightly too large a contribution for JC (£)d£ between 4 = 0 and £ = 0.5M. We will assume that subsequent positive and negative parts of C (£) will cancel each other and contribute nothing to J*C (£)d£ from £=0.5Mto£=R. In the example being discussed, the value of rjhe correlation length, a = 25 cm. Having found a suitable way to determine the equivalent correlation length, all of our algebra for the sound fluctuations due to a Gaussian correlation func- tion is assumed to be applicable, even though the true correlation function is by no means Gaussian. The general method for the determination of the correlation length is found by referring to Fig. 6. The correlation function is down to e at x £ = x t a = 1.8 (3.39) therefore for our approximation 1.8 a = xt Furthermore, using the empirical approximation to the transition wave number that we developed from Fig. 3, Eq. (2.8), we have the generalization for the upper ocean. a = 0.2eH/4° where H = depth, m (3.40) a = temperature (or index) correlation distance in m. It is evident from the data scatter of Fig. 3, from which the constant B and H are determined, that (3.40) can be only a guideline for the prediction of the correlation distance. 34 Although a often decreases with increasing depth, a general rule describing (j as a function of depth is not possible at this time. In fact, when such a rela- tion is found there will certainly be other important parameters in the equation, such as time of day, to which a is particularly sensitive at shallow depths. When a is measured at the place and time of an acoustics experiment, Eq. (3.40) provides a value of correlation length, a, that completes the needs for evaluating q or a . For example, Shvachko (1967) has made measurements of a and CAV af depMs 20, 35, 40M using sound of frequency 25 kHz. If we take Schvachko's value of a / R/ k/ and use his depth in Eq. (3.40) to obtain a, we can calculate q from (3.17) and predict his CAV from (3.37) or the graph Fig. 8. This has been done in the following way: ILLUSTRATIVE EXAMPLE Shvachko (1967) has performed an acoustic fluctuation experiment at frequency 25 kHz, range 20 m, in the sea of Norway. The medium had a standard deviation of the excess index of refraction n = 17.8 x 10 . We are to calculate the predicted cr . At b. /depth 20 m reading from Fig. 3 x = 0.055 cm"" or using (2 . 8) x = 9.0e~ = 5.45 m~ 1 8 therefore from (3.39) a = — '— = 0.33 m t 20/40 or directly from (3.40) a = 0.2 e ' = 0.33 m CO 2 (25.x 103) ._. e -1 k= — = -rh lnA 104.6 m c 1.5 x 10s3 calculate D= 7—2 n i. cl, 0?2\ = 84« >>l ka^1 = (104.6)(.33^) therefore, using the Gaussian assumption (3.17) for D »1 A 2 u = (0.89) (17.8 x 10"5)2(104.6)2(240.)(0.33) = 0.0244 oA = 0.16 This is to be compared with Shvachko's CAV = 0.17 35 Since the