Munk1981 Wunsch Chapter9

Embed Size (px)

Citation preview

  • 8/13/2019 Munk1981 Wunsch Chapter9

    1/28

  • 8/13/2019 Munk1981 Wunsch Chapter9

    2/28

    routinely applied in the fields of optics and acoustics),and the spectral representation of surface waves hadproven very useful. It became clear that internal waves,too, occupy a frequency continuum, over some six oc-taves extending from inertial to buoyant frequencies.[The high-frequency cutoff had been made explicit byGroen (1948).]With regard to the vertical modes, thereis sufficient energy in the higher modes that for manypurposes the discrete modal structure can be replacedby an equivalent three-dimensional continuum.We have already referred to the measurements byUfford and by Lafond at horizontally separated points.Simultaneous current measurements at vertically sep-arated points go back to 1930 Ekman and Helland-Hansen, 1931). In all these papers there is an expressionof dismay concerning the lack of resemblance betweenmeasurements at such small spatial separations of os-cillations with such long periods. I believe (from dis-cussions with Ekman in 1949) that this lack of coher-ence was the reason why Ekman postponed for 23 years(until one year before his death) the publication of"Results of a Cruise on Board the 'Armauer Hansen' in1930 under the Leadership of Bjrnm Helland-Hansen"(Ekman, 1953). But the decorrelation distance is justthe reciprocal of the bandwidth; waves separated inwavenumber by more than Ak interfere destructivelyat separations exceeding (Ak)-. The small observedcoherences are simply an indication of a large band-width.

    The search for an analytic spectral model to describethe internal current and temperature fluctuations goesback over many years, prompted by the remarkablesuccess of Phillips's (1958) saturation spectrum for sur-face waves. I shall mention only the work of Murphyand Lord (1965), who mounted temperature sensors inan unmanned submarine at great depth. They foundsome evidence for a spectrum depending on scalarwavenumber as k-5'3, which they interpreted as theinertial subrange of homogeneous, isotropic turbu-lence. But the inertial subrange is probably not appli-cable (except perhaps at very small scales), and thefluctuations are certainly not homogeneous and notisotropic.Briscoe (1975a) has written a very readable accountof developments in the early 1970s. The interpretationof multipoint coherences in terms of bandwidth wasthe key for a model specturm proposed by Garrett andMunk (1972b). The synthesis was purely empirical,apart from being guided by dimensional considerationsand by not violating gross requirements for the finite-ness of certain fundamental physical properties. Sub-sequently, the model served as a convenient "straw-man" for a wide variety of moored, towed and"dropped" experiments, and had to be promptly mod-ified [Garrett and Munk (1975),which became known

    as GM75 in the spirit of planned obsolescence]. Therehave been further modifications [see a review paper byGarrett and Munk (1979)]; the most recent version issummarized at the end of this chapter.

    The best modem accounts on internal waves are byO. M. Phillips '(1966b), Phillips (1977a), and Turner(1973a).Present views of the time and space scales ofinternal waves are based largely on densely sampledmoored, towed, and dropped measurements. The pi-oneeringwork with mooringswas done at site D in thewestern North Atlantic (Fofonoff, 1969; Webster,1968). Horizontal tows of suspended thermistor chains(Lafond, 1963; Charnock, 1965)were followed by towedand self-propelled isotherm-following "fishes" (Katz,1973; McKean and Ewart, 1974). Techniques fordropped measurements were developed along a numberof lines: rapidly repeated soundings from the stableplatform FLIP by Pinkel (1975), vertical profiling ofcurrents from free-fall instruments by Sanford (1975)and Sanford, Drever, and Dunlap (1978), and verticalprofiling of temperature from a self-contained yo-yoingcapsule by Cairns and Williams (1976). The three-di-mensional IWEX (internal wave experiment) array isthe most ambitious to date (Briscoe, 1975b). These ex-periments have served to determine selected parame-ters of model spectra; none of them so far, not evenIWEX, has been sufficiently complete for a straight-forward and unambiguous transform into the multi-dimensional (co,k)-spectrum. The FLIP measurementscome closest, giving an objective spectrum in the twodimensions o, k, with fragmentary information on ka,k,. Otherwise only one-dimensional spectra can beevaluated from any single experiment, and one is backto model testing. Yet in spite of these observationalshortcomings, there is now evidence for some degreeof universality of internal wave spectra, suggesting thatthese spectra may be shaped by a saturation process(the interior equivalent of whitecaps), rather than byexternal generation processes.Internal waves have surface manifestations consist-ing of alternate bands of roughened and smooth water(Ewing, 1950; Hughes, 1978), and these appear to bevisible from satellites (figure 9.1).High-frequency sonarbeams are a powerful tool for measuring internal waverelated processes in the upper oceans (figures 9.2, 9.3).The probing of the deep ocean interior by acoustics isultimately limited by scintillations. due to internalwaves (Flatte et al., 1979; Munk and Wunsch, 1979)just as the "diffraction-limited" telescope has its di-mensions set by the small-scale variability in the upperatmosphere.It will be seen that internal waves are a lively subject.The key is to find the connections between internalwaves and other ocean processes. The discovery of everfiner scales, down to the scale of molecular processes,

    265Internal Waves

  • 8/13/2019 Munk1981 Wunsch Chapter9

    3/28

    Figure 9.I SEASATsynthetic aperture radar image off Cabo area. The pattern in the right top area is most likely formedSan Lzaro, Baja California (2448'N, 11218'W) taken on 7 by internal waves coming into the 50 fathom line. (I amJuly 1978. Scale of image nearly matches that of bathymetric indebted to R. Bernstein for this figure.)

    CT UT23 24 25 23 24I i1 -_ I 1 I I 1

    I I I

    (30 -

    Ar

    on

    AetC . ' ' '-ae*_ ~ ~ ~ ~ : 4_..'_ " 'e ' n,.: 7'

    ,,-:t:? ,) : . :. ,,. ~, .?,0 5 10Time (min)Figure 9.2 The water column is insonified with a narrowdownward sonar beam of 200 kHz (wavelength 0.75 cm). Thedark band is presumably a back-scattering layer convolutedby shear instabilities. In a number of places the instabilitieshave created density inversions. This is confirmed by the twoo-t-profiles. The acoustic reflection from the sinking CTDalong the steeply slanting lines shows the depth-time historyof the rr,-profiles.The profiling sound source was suspendedfrom a drifting ship. The horizontal distance between over-turning events was estimated to be 60-70 m. (I am indebtedto Marshall Orr of Woods Hole Oceanographic Institution forthis figure; see Haury, Briscoe, and Orr, 1979.)

    266Walter Munk

    0

    15Ev:

    : ..ti,A h Io, ' '_

  • 8/13/2019 Munk1981 Wunsch Chapter9

    4/28

    has been a continuing surprise to the oceanographiccommunity for 40 years. Classical hydrographic castsemployed reversing (Nansen) bottles typically at 100-m intervals in the upper oceans beneath the thermo-cline, and half-kilometer intervals at abyssal depths.Only the gross features can be so resolved. Modemsounding instruments (BT, STD, CTD) demonstrateda temperature and salinity3 fine structure down to me-ter scales. An early clue to microstructure was thesteppy traces on the smoked slides of bathythermo-graphs. These steps were usually attributed to "stylusstiction," and the instruments suitably repaired.Free-fall apparatus sinking slowly (-0.1 ms-}) andemploying small, rapid-response (-0.01 s) transducers,subsequently resolved the structure down to centime-ter scales and beyond. The evolving terminologygross structure:fine structure:microstructure:

    a,CW

    larger than 100 m vertical1 m to 100 m verticalless than 1 m vertical 01 02 03 04II June 1977 (hr)is then largely based on what could be resolved in agiven epoch (see chapter 14). The fine-structure meas-urements of temperature and salinity owe much oftheir success to the evolution of the CTD (Brown,1974). The pioneering microscale measurements weredone by Woods (1968a)and by Cox and his collaborators(Gregg and Cox, 1972; Osborn and Cox, 1972). Meas-urements of velocity fine structure down to a few me-ters have been accomplished by Sanford (1975) andSanford, Drever, and Dunlap (1978). Osborn (1974,1980)has resolved the velocity microstructure between40 and 4 cm. Evidently velocity and temperature struc-ture have now been adequately resolved right down tothe scales for which molecular processes become dom-inant. At these scales the dissipation of energy andmean-square temperature gradients is directly propor-tional to the molecular coefficients of viscosity andthermal diffusivity. The dissipation scale for salinityis even smaller (the haline diffusivity is much smallerthan the thermal diffusivity) and has not been ade-quately resolved. The time is drawing near when weshall record the entire fine structure and microstruc-ture scales of temperature, salinity and currents [andhence of the buoyancy frequency N(z) and of Richard-son number Ri(z)] from a single free-fall apparatus.Perhaps the discovery of very fine scales couldhave been anticipated. There is an overall ocean bal-ance between the generation and dissipation of mean-square gradients. Eckart (1948) refers to the balancingprocesses as stirring and mixing. Garrett (1979) has putit succinctly: "Fluctuations in ocean temperature pro-duced by surface heating and cooling, and in salinitydue to evaporation, precipitation, run-off and freezing,are stirred into the ocean by permanent current sys-tems and large scale eddies." Mixing ultimately occurs

    Figure 9.3 Measurements of Doppler vs. range were made at2-minute intervals with a quasi-horizontal 88-kHz soundbeam mounted on FLIP at a depth of 87 m. Bands of alternat-ing positive and negative Doppler in velocity contours) arethe result of back scatter from particles drifting toward andaway from the sound source (the mean drift has been re-moved). The velocities are almost certainly associated withinternal wave-orbital motion. The range-rate of positive ornegative bands gives the appropriate projection of phase ve-locity. The measurements are somewhat equivalent to suc-cessive horizontal tows at 3000 knots! (I am indebted to Rob-ert Pinkel of Scripps Institution of Oceanography for thisfigure.)through dissipation by "molecular action on small-scale irregularities produced by a variety of processes."The microstructure (where the mean-square gradientslargely reside) are then a vital component of oceandynamics. This leaves open the question whether mix-ing is important throughout the ocean, or whether itis concentrated at ocean boundaries and internal fronts,or in intense currents an extensive discussion may befound in chapter 8).

    What are the connections between internal wavesand small-scale ocean structure? Is internal wavebreaking associated with ocean microstructure? Isthere an associated flux of heat and salt, and hencebuoyancy? Does the presence of internal waves in ashear flow lead to an enhanced momentum flux, whichcan be parameterized in the form of an eddy viscosity?What are the processes of internal-wave generation anddecay? I feel that we are close to having these puzzlesfall into place (recognizing that oceanographic "break-throughs" are apt to take a decade), and I am uncom-fortable with attempting a survey at this time.Forty years ago, internal waves played the role of anattractive nuisance: attractive for their analytical ele-

    267Internal Waves

  • 8/13/2019 Munk1981 Wunsch Chapter9

    5/28

    gance and their accessibility to a variety of experimen-tal methods, a nuisance for their interference withwhat was then considered the principal task of physicaloceanography, namely, charting the "mean" densityfield. Twenty years from now I expect that internalwaves will be recognized as being intimately involvedwith the vertical fluxes of heat, salt, and momentum,and so to provide a vital link in the understanding ofthe mean fields of mass and motion in the oceans.9.1.1 Preview of This ChapterWe start with the traditional case of a two-layer ocean,followed by a discussion of continuous stratification:constant buoyancy frequencyN, N decreasing withdepth, a maximum N (thermocline), a double maxi-mum. Conditions are greatly altered in the presence ofquite moderate current shears. Short (compliant) inter-nal waves have phase velocities that are generallyslower than the orbital currents associated with thelong (intrinsic) internal waves, and thus are subject tocritical layer processes. There is further nonlinear cou-pling by various resonant interactions.

    Ocean fine structure is usually the result of internal-wave straining, but in some regions the fine structureis dominated by intrusive processes. Microstructure isconcentrated in patches and may be the residue ofinternal wave breaking. Little is known about thebreaking of internal waves. Evidently, there are twolimiting forms of instability leading to breaking: ad-vective instability and shear instability.The chapter ends with an attempt to estimate theprobability of wave breaking, and of the gross verticalmixing and energy dissipation associated with thesehighly intermittent events. An important fact is thatthe Richardson number associated with the internalwave field is of order 1. Similarly the wave field iswithin a small numerical factor of advective instabil-ity. Doubling the mean internal wave energy can leadto a large increase in the occurrence of breaking events;halving the wave energy could reduce the probabilityof breaking to very low levels. This would have theeffect of maintaining the energy level of internal waveswithin narrow limits, as observed. But the analysis isbased on some questionable assumptions, and the prin-cipal message is that we do not understand the prob-lem.9.2 Layered OceanWe start with the conventional discussion of internalwaves at the boundary between two fluids of differentdensity. The configuration has perhaps some applica-tion to the problem of long internal waves in the ther-mocline, and of short internal waves in a stepwise finestructure.

    Following Phillips (1977a), this can be treated as alimiting case of a density transition from p, above z =-h to pi beneath z = -h, with a transition thickness8h (figure 9.4). The vertical displacement {(z) as a peakat the transition, and the horizontal velocity u z)changes sign, forming a discontinuity (vortex sheet) inthe limit 8h -O0. For the second mode (not shown), C(z)changes sign within the transition layer and u(z)changes sign twice; this becomes unphysical in thelimit Ah - 0. For higher modes the discontinuities areeven more pathological, and so a two-layer ocean isassociated with only the gravest internal mode.For the subsequent discussion it is helpful to give asketch of how the dependent variables are usually de-rived and related. The unknowns are u,v,w,p (aftereliminating the density perturbation), where p is thedeparture from hydrostatic pressure. The four un-knowns are determined by the equations of motion andcontinuity (assuming incompressibility). The linear-ized x,y equations of motion are written in the tradi-tional f-plane; for the vertical equation it is now stan-dard [since the work of Eckart (1960)] to display thedensity stratification in terms of the buoyancy (orBrunt-Vaisala) frequency

    (9.1)thus givingOw = 1 P- N 2 = 0.Ot po tThe last term will be recognized as the buoyancy force-g 8ppo of a particle displaced upwards by an amountl = fwdt.

    For propagating waves of the form (z) expi(kx - t)the equations can beand 5.7) into combined (Phillips, 1977a, 5.2

    o U- U -o o

    'l

    p N-_Z

    + ' l 0iif~

    iH

    Figure 9.4 A sharp density transition from p, to PI akes placebetween the depths -z = h - 8h and -z = h + 18h.This isassociated with a delta-like peak in buoyancy frequency N(z).Amplitudes of vertical displacement 4(z), horizontal velocityu(z), and shear u'(z) = duldz are sketched for the gravestinternal wave mode.

    268Walter Munk

    -g['--[~(d~jz iabaticl

  • 8/13/2019 Munk1981 Wunsch Chapter9

    6/28

    d2--+ Nk2 - =o. (9.2)dz2 w 2 -f 2The linearized boundary conditions are 00 at thesurface and bottom.

    A simple case is that of f = 0 and N = 0 outside thetransition layer. We have thenC.=A sinhkz,1 = B sinhk(z + H),above and below the transition layer, respectively. The

    constants A and B are determined by patching the ver-tical displacement at the transition layer:u = =a at z = -h.

    The dispersion relation is found by integrating Eq. (9.2)across the transition layer:C 1= -k24 N(Z) - W2

    - 9(g- p - 28h) at z = -h,

    transition where N reaches a maximum, but just theopposite is true. To prove this, we use the condition ofincompressibility, iku(z) - io4' = 0, and equation (9.2)to obtain

    U= c = N 2z) - kak o~~~~c (9.3)and so u' - N2 for small wIN; accordingly u'/N variesas N. Thus the layers of largest gravitational stability(largest N) are also the layers of largest shear instability(largest u'/NI).9.3 Continuously Stratified OceanThe simplest case is that of constant N. The solutionto (9.2) is4(z)= a sinmz, m2 = k2N ' -(a _2 2 (9.4)with m so chosen that ; vanishes at z = -H. Solvingfor to2,

    where 5' ddz. In the limit of small k h, that is, forwaves long compared to the transition thickness, theforegoing equations lead to the dispersion relation

    g(2plp)kcothkh cothk(Hh)For a lower layer that is deep as compared to a wave-length, the denominator becomes cothk(h - H) + 1.If the upper layer is also deep, it becomes 1 + 1, and

    1 pk P lk P'- P,2 p 2 g(pt + p.)'As p, - 0, O --) gk, which is the familiar expressionfor surface waves in deep water.The case of principal interest here is that of an iso-lated density transition 8p

  • 8/13/2019 Munk1981 Wunsch Chapter9

    7/28

    N _U

    -z

    -H

    C (z)

    Figure 9.6 Vertical displacements i(z) n a constant-N ocean,for modes = 1 and i = 3.

    z near inertialfrequency

    kI>8x

    Cg

    z near buoyancyfrequency

    i,kCg

    Figure 9.7 The wavenumber vector k = (k,m) and group ve-locity cg near the inertial frequency to = f + E) and near thebuoyancy frequency {to = N - E}, espectively. A packet ofwave energy is projected on the (x,z)-plane. Crests and troughsin the wave packet are in a plane normal to k, and travel withphase velocity c in the direction k. The wave packet travelswith group velocity cg at right angles to k, thus sliding side-ways along the crests and troughs. The particle velocity u (notshown) is in the planes at right angles to k.

    k = (k, m) are easy to derive, but hard to visualize. Con-sider a wave packet (figure 9.7) with crests and troughsalong planes normal to the paper and inclined withrespect to the (x,z)-axis as shown. The phase velocityis in the direction k normal to the crests, but the groupvelocity cg is parallel to the crests, and the wave packetslides sideways. k is inclined to the horizontal bym (N 2 - a1/2tan8 = = (2_f2 (9.6a)

    and so the angle is steep for inertial waves (o = f + E)and flat for buoyancy waves (0o= N - E).The energypacket is propagated horizontally for inertial waves,and vertically for buoyancy waves, but the group ve-locity goes to zero at both limits.The flow u = (u, w) takes place in the plane of thecrest and troughs. For inertial waves, particles move inhorizontal circles. The orbits become increasingly el-liptical with increasing frequency, and for buoyancywaves the particle orbits are linear along the z-axis, inthe direction of cg.The wavenumber k is always nor-mal to both cg and u. [The nonlinear field accelerations(u V)u vanish for an isolated elementary wave train,leading to the curiosity that the linear solution is anexact solution.] Readers who find it difficult to visu-alize (or believe) these geometric relations should referto the beautiful laboratory demonstrations of Mowbrayand Rarity (1967).It is not surprising, then, that internal waves will dounexpected things when reflected from sloping bound-aries. The important property is that the inclination 0relative to the x-axis depends only on frequency [equa-tion (9.6a)]. Since frequency is conserved upon reflec-tion, incident and reflected 0 must be symmetric withrespect to a level surface rather than with respect tothe reflecting surface. At the same time the flow u forthe combined incident and reflected wave must be par-allel to the reflecting boundary. For a given co, here isa special angle for which the orbital flow is parallel tothe boundary. This requires that the boundary be in-clined at a slopetan/i(z) tan(90- =0) [:z - L2] (9.6b)It can be shown that for slopes steeper than /, theenergy of "shoreward" traveling internal waves is re-flected "seaward": for slopes of less than /3, the energyis forward reflected. Repeated reflections in a wedge-shaped region such as the ocean on the continentalslope can lead to an accumulation of energy at eversmaller scales (Wunsch, 1969). For a given slope, wecan expect an amplification of the internal waves atthe frequency codetermined by (9.6b). Wunsch (1972b)has suggested that a peak in the spectrum of temper-ature fluctuations measured southeast of Bermuda

    270Walter Munk

    ! i1

    o0l

  • 8/13/2019 Munk1981 Wunsch Chapter9

    8/28

    could be so explained. Pertinent values are N =2.6 cph, f = 0.045 cph, and i- 13. Equation (9.6b)gives o = 0.59 cph, in agreement with the observedspectral peak at 0.5 cph.9.4 TurningDepths and Turning LatitudesFigure 9.8 shows the situation for an ocean with vari-able N(z). For frequencies that are less than N through-out the water column, the displacements are similar tothose for constant N (figure 9.6) except that the posi-tions of the maxima and zeros are displaced somewhatupward, and that the relative amplitudes are somewhatlarger at depth. The important modification occurs forfrequencies that exceed N(z) somewhere within thewater column. At the depths ZT where o = NIZT), calledturning depths, we have the situation shown to theright in figure 9.8. Equation (9.2) is locally of the formC'+ z = 0 wherez is now a rescaled vertical coordinaterelative to ZT. The solution (called an Airy function)has an inflection point at the turning depth (here z =0), is oscillatory above the turning depth, and is expo-nentially damped beneath. The amplitudes are some-what larger just above the turning depth than at greaterdistance, but nothing very dramatic happens.The refraction of a propagating wave packet is illus-trated in figure 9.9. As the packet moves into depthsof diminishing N(z) the crests and troughs turn steeper,and the direction of energy propagation becomes morenearly vertical. The waves are totally reflected at theturning depth ZT where = N(zT). Modal solutions;(z) x exp i(kx - ot) with 6(z)as illustrated in figure9.8 can be regarded as formed by superposition of prop-agating waves with equal upward and downward en-ergy transport. The wave energy remains trapped be-tween the surface and the turning depth.The common situation for the deep ocean is themain thermocline associated with a maximum in N(z).Internal waves with frequencies less than this maxi-mum are in a waveguide contained between upper andlower turning depths. For relatively high (but stilltrapped) frequencies the sea surface and bottom bound-aries play a negligible role, and the wave solutions canbe written in a simple form (Eriksen, 1978). The bottomboundary condition (9.5) for a constant-N ocean, e.g.,mjH = jr, j = 1,2,..., is replaced in the WKB approxi-mation by

    n0

    -z0

    WI, W

    ~z~I

    C=W,

    Figure 9.8 Vertical displacements (z) in a variable-N ocean,for modes i = 1 and i = 3. o, is taken to be less than N(z) atall depths. o2 is less than N(z) in the upper oceans above z =-z, only.

    -z

    N-N(z)

    Figure 9.9 Propagation of a wave packet in a variable-N(z)ocean without shear (U = constant). The turning depth ZToccurs when w = N(zT).

    u-_z

    k

    . . -

    IN w 24\112 Figure g.rom jb= s l) /2= rN/No, 19.7) oceanwithC(c).where b is a representative thermocline (or stratifica-tion) scale. Equation (9.7) assures an exponential atten-uation outside the waveguide. For the case of a doublepeak in N(z) with maxima N1 and N2, the internal waveenergy is concentrated first at one thermocline, then

    Propagation of a wave packet in a constant-Nshear. The critical depth z, occurs where U =

    27IInternal Waves

    .

    I

  • 8/13/2019 Munk1981 Wunsch Chapter9

    9/28

    the other, migrating up and down with a frequencyIN - N2I (Eckart, 1961).This is similar to the behaviorof two loosely coupled oscillators. The quantum-me-chanical analogy is that of two potential minima andthe penetration of the potential barrier between them.There is a close analogy between the constant- andvariable-N ocean, and the constant- and variable-focean (the f-plane and /3-plane approximations). For afixed , the condition co= f = 2 sin T determines theturning latitude T. Eastward-propagating internalgravity waves have solutions of the formn(y)4(z)expi(kx - cot).The equation governing the localnorth-south variation is (Munk and Phillips, 1968)4' + yr- = 0, " = d2 /qdy,where y is the poleward distance (properly scaled) fromthe turning latitude. This is in close analogy with theup-down variation near the turning depth, which isgoverned by" + z4 = 0, 4"= d24/dz2 .

    Thus il(y)varies from an oscillatory to an exponentiallydamped behavior as one goes poleward across the turn-ing latitude. Poleward-traveling wave packets are re-flected at the turning latitude.From an inspection of figure 9.7, it is seen that theroles of horizontal and vertical displacements are in-

    ()

    ._

    C

    -m.

    terchanged in the N(z) and f(y) turning points. In theN(z) case the motion is purely vertical; in the f(y) casethe motion is purely horizontal (with circular polari-zation).It has already been noted that nothing dramatic isobserved in the spectrum of vertical displacement (orpotential energy) near o = N-only a moderate en-hancement, which can be reconciled to the behavior ofthe Airy function (Desaubies, 1975; Cairns and Wil-liams, 1976). Similarly we might expect only a mod-erate enhancement in the spectrum of horizontal mo-tion (or kinetic energy) near co= f. In fact, the spectrumis observed to peak sharply. If the horizontal motion iswritten as a sum of rotary components (Gonella, 1972),it is found that the peak is associated with negativerotation (clockwise in the northern hemisphere).I have made a parallel derivation of the spectra at thetwo turning points (figure9.11), assuming horizontallyisotropic wave propagation within the entire equatorialwaveguide. It turns out that the buoyancy peak is infact much smaller than the inertial peak at moderatelatitudes. But at very low latitude the inertial peakvanishes. This is in accord with the equatorial obser-vations by Eriksen (1980).Fu (1980)gives an interestingdiscussion of the relative contributions to the spectralpeak at the local inertial frequency co= foca from twoprocesses: (1) local generation of resonant inertial

    N i/2Nmax 2

    .if .2f .5f f 2f 5fwFigure 9.II Enhancement of the kinetic-energy spectrum(left) and of the potential-energy spectrum (right) at the iner-tial and buoyancy frequencies, respectively. The inertial spec-

    .2N .5N N 2N

    trum is drawn for latitudes 1 , 5 , 10 , 30, 45 . The buoyancyspectrum is drawn for two depths, corresponding to N = , 4times the maximum buoyancy frequency.272Walter Munk

  • 8/13/2019 Munk1981 Wunsch Chapter9

    10/28

    waves co = ocal;and (2) remote generation of waves ofthe same frequency o = fiocal t lower latitudes (wheref < focal}. Figure 9.11 is drawn for case 2 under theassumption that the equatorial waveguide is filled withhorizontally isotropic, freely propagating radiation.Take the curve marked 30, say. Then for o > f astation at lat. 30 is within the equatorial waveguide;for co < f the spectrum is the result of evanescentextensions from a waveguide bounded by lower lati-tudes. Over rough topography and in regions of strongsurface forcing, the case can be made for local genera-tion of the inertial peak. It would seem that the buoy-ancy peak at mid-depth must always be associated withremote generation.9.5 ShearInternal waves are greatly modified by an underlyingshear flow.4 A variable U(z) can have a more traumaticeffect on internal waves than a variable Niz). For readycomparison with figure 9.9 showing the effect of avariable N(z) on a traveling wave packet, we havesketched in figure 9.10 the situation for a wave packettraveling in the direction of an increasing U(z). As thewave packet approaches the "critical depth" zc wherethe phase velocity (in a fixed frame of reference) equalsthe mean flow, c = U(zc), the vertical wavenumberincreases without limit (as will be demonstrated).

    For the present purpose we might as well avoid ad-ditional complexities by setting f = 0. The theoreticalstarting point is the replacement of Otby Ot + Ua, +w Oz n the linearized equations of motion. The resultis the Taylor-Goldstein equation [Phillips (1977a, p.248)]:d% N U - =dz (U - C)2 U - c =

    dzdUdz2 I

    -oo0

    -z

    u (z)-0

    0

    C z) ~0

    I+

    Figure 9.I2 First mode vertical displacements (z) in aCouette flow (constant U' and constant N), for U'/N = 0, + 1.Waves move from left to right, and U is positive in the direc-tion of wave propagation. (Thorpe, 1978c.)

    -zU

    -h

    F

    L

    N(z) , U'(z)--0

    0 +

    o

    Do

    I

    Figure 9.I3 Similar to figure 9.12, but with U' and N confinedto a narrow transition layer.

    (9.8)

    where c is the phase velocity in a fixed reference frame.(This reduces tod2 f Nk2_ o 2d2 +k2 = 0dz-- (9.9)for U = 0.) The singularity at the critical depth whereU = c is in contrast with the smooth turning-pointtransition at N = o; this is the analytic manifestationof the relative severity of the effect of a variable U(z)versus that of a variable N(z).Thorpe (1978c) has computed the wave function ;(z)for (1) the case of constant N and U' and (2) the casewhere N and U' are confined to a narrow transitionlayer. The results are shown in figures 9.12 and 9.13.The profiles are noticeably distorted relative to the case

    273Internal Waves

    \

    r

    -H

    > -::

    A

    -- UN

  • 8/13/2019 Munk1981 Wunsch Chapter9

    11/28

    of zero shear, with the largest amplitudes displacedtoward the level at which the mean speed (in the di-rection of wave propagation) is the greatest. Finite-am-plitude waves have been examined for the case 2.Where there is a forward5 flow in the upper level (in-cluding the limiting case of zero flow), the waves havenarrow crests and flat troughs, like surface waves; withbackward flow in the upper layer, the waves have flatcrests and narrow troughs. Wave breaking is discussedlater.9.5.1 Critical Layer Processes 6The pioneering work is by Bretherton (1966c), and byBooker and Bretherton (1967). Critical layers have beenassociated with the occurrence of clear-air turbulence;their possible role with regard to internal waves in theoceans has not been given adequate attention.Following Phillips (1977a), letohF =kU + w, co k= (k2 +m 2 )1 2 = cosdesignate the frequency in a fixed reference frame. U(z)is the mean current relative to this fixed frame, andtoF - kU = o is the intrinsic frequency [as in (9.5)], asit would be measured from a reference frame driftingwith the mean current U(z).Bretherton (1966c) has given the WKB solutions forwaves in an ocean of constant N and slowly varyingU. [It is important to note the simplification to (9.8)when U" = 0 at the critical layer.] Near the criticiallayer depth z,, the magnitudes of w, u, and of the ver-tical displacement vary asw - IZ - zCl'"', u -IZ - ZC{1/2,

    The quantities OFand k are constant in this problem,but m and ot are not. The vertical wavenumber in-creases, whereas the intrinsic frequency decreases as awave packet approaches its critical layer:m - Iz - zc,-', - Iz - Z.A sketch of the trajectory is given in figure 9.10. Wavesare refracted by the shear and develop large verticaldisplacements (even though w -- 0), large horizontalvelocities u, and very large induced vertical shears u'.This has implications for the dissipation and breakingof internal waves.

    For Ri > , Booker and Bretherton (1967) derived anenergy transmission coefficientp = exp(-27rRi - ). (9.11)In the usual case, U' > and p issmall. This is interpreted as wave energy and momen-tum being absorbed by the mean flow at z,. As Ri -+o, p -- 0, consistent with the WKB prediction of Breth-erton (1966c) that a wave packet approaches but neverreaches the critical layer.

    The small coefficient of transmission for Richardsonnumbers commonly found in the ocean implies thatthe critical layer inhibits the vertical transfer of waveenergy. This effect has been verified in the laboratoryexperiments of Bretherton, Hazel, Thorpe, and Wood(1967). When rotation is introduced, the energy andmomentum delivered to the mean flow may altema-tively be transferred from high-frequency to low-fre-quency waves (if the time scales are appropriate). Thusit is possible that some sort of pumping mechanismmay exist for getting energy into, for example, the high-mode, quasi-inertial internal waves. This mechanismcan be compared with McComas and Bretherton's(1977) parametric instability, a weakly nonlinear inter-action (section 9.6).The work of Bretherton and of Booker and Brethertonhas prompted a great number of critical-layer studies.One of the most interesting extensions was done byJones (1968). Whereas Booker and Bretherton found thecritical layer to be an absorber, not a reflector, whenRi > , Jones found that reflection from the criticallayer is possible when Ri < ; in fact, the reflectedwave amplitude can exceed that of the incident wave.Jones called these waves "overreflected," their energybeing enhanced at the expense of the mean flow. Thisis illustrated in figure 9.14, based on a solution for ahyperbolic-tangent profile intended to display the re-sults of linear theory. Transmission and reflection ra-tios at z = + were derived using definitions of waveenergy density appropriate to moving media. "Over-transmission" as well as overreflection occurs at verysmall Richardson numbers, with the internal wavesgaining energy from the mean flow on both counts.We shall now consider the condition for critical-layerabsorption. Let co,designate the intrinsic frequency ofa wave packet at some depth z, with a mean flow U,in the direction of wave propagation. According to(9.10),

    )F = k U + 1.Let U increase to some value U2 at z2. Then since Fand k are conserved along the trajectory of the wavepacket,cOF= kU 2 + 02For the special case that z 2 is to be a critical depth, wehave J)F kU 2, hence o2 = 0, and soo/lk = U 2 - U.

    The vertical wavenumber of internal waves is givenby the dispersion relation(N2 - c2) 1/2m=k ( 2 f2

    kNo for f

  • 8/13/2019 Munk1981 Wunsch Chapter9

    12/28

    for critical layer processes as rms u from the internalwaves themselves. The internal wave spectrum is thendivided into two parts: (1) the intrinsic part m < mc,which contains most of the energy, and (2) the com-pliant part m > m,, which is greatly modified by in-teraction with the intrinsic flow field. The phase speedfor critical reflection is

    (9.14)C = rmsu,and the critical wavenumber ismc = N/rmsu. (9.15)

    RiFigure 9.I4 Fractional internal wave energy reflected andtransmitted through a mean shear flow U = Utanh(z/d) atconstant N, as a function of the minimum Richardson numberRi = Nd2/U2o. Internal wave energy is lost to the mean flowfor RI2 + TI < 1, orRi > O.:L8;nternal wave energy is gainedfrom the mean flow for R < 0.18. The plot is drawn forRi = 2a2, where a = kd is the dimensionless horizontal wave-number. This corresponds to a wave packet traveling at aninclination of 45 at z = o-. (Ri = a2 corresponds to thelimiting case of vertical group velocity to +x.) (I am indebtedto D. Broutman for this figure.)

    There is the separate question whether the internalwaves at the critical layer will be underreflected, justreflected, or overreflected, and this depends on the am-bient Richardson number. In the underreflected casethere is a flux of energy from the compliant to theintrinsic waves. In the overreflected case the flow isthe other way. For an equilibrium configuration, onemay want to look for a transmission coefficient p nearunity, and the exponential behavior of p(Ri) will thenset narrow bounds to the ambient spectrum. But thisgets us into deep speculation, and had better be left tothe end of this chapter.9.6 Resonant Interactions

    For critical absorption within the interval Az overwhich the mean flow varies by AU, we replace co/kbyAU, and obtain the critical vertical wavenumbermc = N/A U. (9.13)

    R. Weller (personal communication) has analyzed amonth of current measurements off California for theexpected difference A U =- U2 - U,1 in a velocity com-ponent (either of the two components) at two levelsseparated by Az = z2 - z,. The observations are, ofcourse, widely scattered, but the following values giverepresentative magnitudes:Az in mAU in cm s-1 0 10 25 500 4 7' 10

    0 7 10 12100

    15 (upper 100 m)15 (100-300 m depth)For AU = 10 cms - 1 and N = 0.01 s- (6 cph), (9.13)gives mc = 10-3 cm-l (16 cpkm). Internal waves withvertical wavelengths of less than 60 m are subject tocritical-layer interactions;.A large fraction of the measured velocity differenceAU can be ascribed to the flow field u (z) of the internalwaves themselves, and deduced from the model spec-tra. The expected velocity difference increases to /2times the rms value as the separation increases to thevertical coherence scale, which is of order 100 m. Heremost of the contribution comes from low frequenciesand low wavenumbers. I am tempted to interpret AU

    Up to this point the only interactions considered arethose associated with critical layers. In the literaturethe focus has been on the resonant interaction of wavetriads, using linearized perturbation theory. There aretwo ways in which critical layer interactions differfrom resonant interactions: (1) compliant waves ofanywavenumber and any frequency are modified, as longas c equals u somewhere in the water column; and (2)the modification is apt to be large (the ratio u/c beinga very measure of nonlinearity). For the wave triads,the interaction is (1) limited to specific wavenumbersand frequencies, and (2) assumed to be small in theperturbation treatment. 7 To borrow some words ofO. M. Phillips (1966b), the contrast is between the"strong, promiscuous interactions" in the critical layerand the "weak, selective interactions" of the triads.The conditions for resonance arek l + k2 = k, 01 + (02 = (03,where ki = (k,li, mi), and all frequencies satisfy thedispersion relation wco(k,). esonant interactions arewell demonstrated in laboratory experiments. For atransition layer (as in figure 9.4), Davis and Acrivos(1967) have found that a first-order propagating mode,which alternately raises and lowers the transitionlayer, was unstable to resonant interactions, leading toa rapid growth of a second-order mode, which alter-nately thickens and thins the transition layer like a

    275Internal Waves

  • 8/13/2019 Munk1981 Wunsch Chapter9

    13/28

    propagating link sausage. Martin, Simmons, andWunsch (1972) have demonstrated a variety of resonanttriads for a constant-N stratification.

    Among the infinity of possible resonant interactions,McComas and Bretherton (1977) have been able toidentify three distinct classes that dominate the com-puted energy transfer under typical ocean conditions.Figure 9.15 shows the interacting propagation vectorsin (k,m )-space. The associated frequencies o areuniquely determined by the tilt of the vectors, in ac-cordance with (9.4). Inertial frequencies (between f and2f, say) correspond to very steep vectors, buoyancy fre-quencies (between 2Nand N) to flat vectors, as shown.Elastic scattering tends to equalize upward anddownward energy fluxes for all but inertial frequencies.Suppose that k3 is associated with waves generatednear the sea surface propagating energy downward (atright angles to k3, as in figure 9.7). These are scatteredinto k,, with the property m, = -m3, until the upwardenergy flux associated with k, balances the downwardflux by k3. The interaction involves a near-intertialwave k2 with the property m2 2m3. (The reader willbe reminded of Bragg scattering from waves having halfthe wavelength of the incident and back-scattered ra-diation.) Similarly, for bottom-generated k, waves withupward energy fluxes, elastic scattering will transferenergy into k3 waves.Induced diffusion tends to fill in any sharp cutoffsat high wavenumber. The interaction is between twoneighboring wave vectors of high wavenumber and fre-quency, k, and k3, and a low-frequency low-wavenum-ber vector k2 . Suppose the k2 waves are highly ener-getic, and that the wave spectrum drops sharply forwavenumbers just exceeding Ik3 l, such as Ikll. This in-teraction leads to a diffusion of action (energy/w) intothe low region beyond k31, thus causing k, to grow atthe expense of k2.Parametricsubharmonicinstability transfers energyfrom low wavenumbers k2 to high wavenumbers k, ofhalf the frequency, co = 2-, ultimately pushing energyinto the inertial band at high vertical wavenumber.The interaction involves two waves k, and k3 of nearlyopposite wavenumbers and nearly equal frequencies.The periodic tilting of the isopycnals by k2 varies thebuoyancy frequency at twice the frequency of ki andk3 . (The reader will be reminded of the response of apendulum whose support is vertically oscillated attwice the natural frequency.)The relaxation (or interaction) time is the ratio ofthe energy density at a particular wavenumber to thenet energy flux to (or from) this wavenumber. Theresult depends, therefore, on the assumed spectrum.For representative ocean conditions, McComas (inpreparation) finds the relaxation time for elastic scat-tering to be extremely short, of the order of a period,and so up- and downgoing energy flux should be in

    balance. This result does not apply to inertial frequen-cies, consistent with observations by Leaman and San-ford (1975) of a downward flux at these frequencies.The relaxation time for induced diffusion is typicallya fraction of a period! (This is beyond the assumptionof the perturbation treatment.) Any spectral bump isquickly wiped out. The conclusion is that the resonantinteractions impose strong restraints on the possibleshapes of stable spectra.

    In a challenging paper, Cox and Johnson (1979)have drawn a distinction between radiative and dif-fusive transports of internal wave energy. In the ex-amples cited so far, energy in wave packets is radiatedat group velocity in the direction of the group velocity.But suppose that wave-wave interactions randomizethe direction of the group velocity. Then eventuallythe wave energy is spread by diffusion rather than ra-diation. The relevant diffusivity is K = (c), where ris the relaxation time of the nonlinear interactions.Cox and Johnson have estimated energy diffusivitiesand momentum diffusivities (viscosities); they findthat beyond 100 km from a source, diffusive spreadingis apt to dominate over radiative spreading. There is aninteresting analogy to crystals, where it is known thatenergy associated with thermal agitation is spread bydiffusion rather than by radiation. The explanation liesin the anharmonic restoring forces between molecules,which bring about wave-wave scattering at room tem-peratures with relaxation times in the nanoseconds.9.7 BreakingThis is the most important and least understood aspectof our survey. Longuet-Higgins has mounted a broadlybased fundamental attack on the dynamics of breakingsurface waves, starting with Longuet-Higgins and Fox(1977), and this will yield some insight into the inter-nal-wave problems. At the present time we depend onlaboratory experiments with the interpretation of theresults sometimes aided by theoretical considerations.Figure 9.16 is a cartoon of the various stages in anexperiment performed by Thorpe (1978b). A densitytransition layer is established in a long rectangulartube. An internal wave maker generates waves of thefirst vertical mode. Before the waves have reached thefar end of the tube, the tube is tilted through a smallangle to induce a slowly accelerating shear flow. Theunderlying profiles of density, shear, and vertical dis-placement correspond roughly to the situation in figure9.13.For relatively steep waves in a weak positive8 shear,the waves have sharpened crests. At the position of thecrest, the density profile has been translated upwardand steepened (B1). There is significant wave energyloss in this development (Thorpe, 1978c, figure 10).

    276Walter Munk

  • 8/13/2019 Munk1981 Wunsch Chapter9

    14/28

    elastic scattering induced diffusion parametric subharmonicinstability

    kFigure 9.I5 Resonant triads for three limiting classes of in-teraction, according to McComas and Bretherton (1977). Thepropagation vectors are drawn in (k,m)-space. Radial lines

    C(z)-

    designate the tilt of the k vectors for w = f, 2f, IN, n. takingN = 24f.

    AA I

    A

    AtA P,

    Figure 9.i6 Cartoon for various stages of Thorpe's experi-ment. The early stages lead to the development of advectiveinstability (upper three sketches), and the final stage to shearinstability (bottom). Waves are traveling from left to right;the mean flow is forward in the direction of wave propaga-

    tion) above the density transition layer and backward belowthe transition layer. The density profiles along the indicatedvertical sections are shown to the left; a velocity profile isshown to the top right (thin lines give the undisturbed pro-files).

    277Internal Waves

    m

    -------

    -? 1~~~~~~~~~~~~~' -

    i.

  • 8/13/2019 Munk1981 Wunsch Chapter9

    15/28

    With increased positive shear, or with increased time,the particles at the crest accelerate, the isopycnal wavefront becomes momentarily vertical, and a jet of fluidmoves forward of the crest (B2). The resulting densityinversion gives rise to a Rayleigh-Taylor instability,forming a turbulent patch (shaded)whose turbulentenergy is irretrievably lost to the organized wave mo-tion. The turbulent patch becomes fairly well mixed,and introduces a steplike feature into the density pro-file (B3). The patch spreads horizontally under the in-fluence of the ambient stratification, forming blini, orpancakes. The detailed dynamics are complicated (Bar-enblatt and Monin, 1979); it is possible that in theoceans the spreading of the patches is eventually re-tarded by geostrophic confinement.In Thorpe's laboratory experiment, the later stagesof horizontal spreading are interrupted by the suddenformation of billows that grow rapidly, extracting en-ergy from the mean shear flow (bottom of figure). Theirwave length is quite short, only several times thethickness of the transition layer.Hence, Thorpe (1978b, 1979) distinguishes betweentwo types of instability leading to internal wave break-ing. In the case of advective instability, breaking growsout of existing large-amplitude internal waves: moreprecisely, waves associated with steep isopycnal slopes.Eventually the particles in the crest are advected for-ward of the crest, leading to a local density inversionwith the potential for a Rayleigh-Taylor instability.Advective instability can take place in the absence ofambient shear, though it is advanced by shear. The

    a)W00Ua3:

    second type is induced shear instability (Kelvin-Helm-holtz instability in the limit of an abrupt density tran-sition), and can take place even in the absence of any(finite) wave disturbance, but is catalyzed by an exist-ing wave background.The two types appear as end points on a stabilitycurve in slope-shear space, constructed by Thorpe(1978b, 1979) from theory and experiment (figure9.17). Under the conditions described by the author,internal waves on a transition layer are unstable if theirslope exceeds 0.34 in the absence of ambient shear, andif the shear exceeds 2N in the absence of slope. Awayfrom the end points, there is advective instability mod-ified by shear, and shear (K-H) instability modified byadvection. The stability curve for the transition profileis not symmetric, implying that (under the prescribedgeometry) negative shear delays instability.The essential feature of advective instability is thatthe particle speed at the crest eventually exceeds thewave speed. The stability curve in figure 9.16 has beenconstructed from Ucrest = c (carrying the theory to thirdorder in wave slope). This is in fair agreement withexperiment. From a similar point of view, Orlanski andBryan (1969) had previously derived the required criti-cal amplitude for advective instability in the oceans,and have checked their analysis with numerical exper-iments. They conclude that more than enough internalwave energy exists for this type of instability to occur.They also conclude that conditions favor advectiveinstability over shear instability, by the following verysimple argument. From (9.12),

    'Iduced

    Couetteprofile trans 4 L0(nU)03:

    UNSTABLE UNSTABLEI

    2 -2U /N

    Figure 9.I7 Stability diagram for internal waves in a shearflow for a Couette profile (asin figure 9.12) and for a transitionlayer as in figure 9.13). Slope is defined as r waveheight/wavelength. The ordinate is U'/N (or Ri-12 ), with Uin the direction of wave propagation. Forthe transition profile,

    layer

    UNSTABLE

    (U/ N)mxU' and N2 are proportional to sech2(z - h), and henceU'/N - sech(z - h) has its maximum value at the transitionz = h. The curves are drawn for the specific dimensionsdescribed by Thorpe (1978a, 1979).

    278Walter Munk

    -

  • 8/13/2019 Munk1981 Wunsch Chapter9

    16/28

    mu u (1 - co2/N2~ 12N co/k 1 f2/w2 )u

    co/k for f

  • 8/13/2019 Munk1981 Wunsch Chapter9

    17/28

    C

    100

    Er 150O.a200

    200

    MR7(cph)2 C/m

    34 35

    C

    34 35

    Figure 9.18 The water structure at two stations 60 km south-'west of Cabo San Lucas (the southern tip of Baja California),and the associated T-S diagrams (from Gregg, 1975). Themeasurements have been processed to give the fine structureof S, T, a-,, N and the microstructure of dTldz. Note differ-ences in scale. The cuspy T-S diagram for MR7 is an indica-tion of intrusive fine structure.

    280Walter Munk

    MR6(cph)2 CC/m

    :2.L)

    OC

    1

    1

    1

  • 8/13/2019 Munk1981 Wunsch Chapter9

    18/28

    Table 9.1 Variances and Spectra of Vertical Gradients in the Ocean Fine Structure and Microstructure at Two Stations offCabo San Lucas (MR7 and MR6) and in the Mid-Gyre of the Central North Pacific (MSR4)MR7 MR6 MSR4aStrongly intrusive Nonintrusive

    (zT - f C)2)n In 26 9x 10- 1x 10-3(aT) 2 in ) 4x 10- 3 8x10- 4 6x 10- 4Thermal diffusivityin cm2 s - 1 8 b 0.015 0.002Cycles per meter 0.1 1 10 0.1 1 10 0.1 1 10Spectrum of

    zT in ()CPm 2x10 -1 lxlO- 2 xlO- 2 2x10-3 2x10- 4 7x10-5 7x10-4 lx10- 4 3xlO - 50aS in (-m/cpm 2x 10-2 lx10-3 8 x10- 4 5x10- xlO 10 - 5 ? 4x10 - 8x10- 2x 10-5ap in(gms cpm 2x10-9 7x10- 1 7x10-10 2x10-1 2x10-11 ? 2x10-" 1lx10-" 9x10- 12

    Spectrum ofa aTin gcm cpm 2x10- 8 1xlO-10 9 xlO- 9 2x10-10 2x10- 11 ? 3x10- 4x10- 12 lx10- 2baS in V_ m -)cpm 1x10-8 0.6x10-9 0.5x10-9 0.3x10- 10 0.6x10-11 ? 0.3x10-1 5x10- 2 13x10 -1

    p-ing r \2cpm 0.2x10-8 0.7x10- 9 0.7x10-9 2x10-10 2x10-11 ? 2x10-1 1 0x10-12 9x10-' 2a. MSR4 is not necessarily representative for the mid-gyre; subsequent cruises have given larger mean-square gradients.b. The vertical heat flux for MR7 can probably not be modeled by an eddy coefficient (Gregg, 1975).

    and nonintrusive situations, respectively. The conclu-sions are: (1) The ratio of the mean-square gradient tothe mean gradient squared (the "Cox number") for tem-perature is highly variable, from 5000 at MR7 to 2 inthe mid-gyre. Under certain assumptions (Osborn andCox, 1972), the eddy diffusivity is the molecular dif-fusivity times this ratio, giving values all the way from8 to 0.002 cm2 s- ' (but see the footnote to table 9.1).The canonical value of 1 cm2 s- 1 [for which I am partlyresponsible (Munk, 1966)] is of no use locally. (2)Spectral levels in vertical gradients diminish with in-creasing vertical wavenumber up to 1 cpm, and thenlevel off. (3) The relative contributions to the density-gradient spectrum has been estimated from azp =-a T + b ,S, with a = 1.7 x 10- 4 gcm-3 (C)-1, b =8 .x 10- 4 gcm-3(0 o)-1. For MR7 at 0.1 cpm, the meas-ured density gradient is much smaller than that in-ferred from either temperature alone or salinity alone.This is consistent with the near cancellation betweentemperature and salinity for intrusive features. (4) Athigher wavenumbers for MR7, and at all wavenumbersof MR6 and MSR4, the density gradient spectrum is ofthe same order as that inferred from temperature orsalinity alone, thus implying the dominance of internalwaves.

    Probability densities of the temperature gradients arehighly non-Gaussian with an enormous flatness factor(138 for MR7, 55 for MR6) attesting to the patchiness(Gregg, 1975).The construction of meaningful ensem-ble averages in a highly intermittent environment(space and time) is an important task for the future.To return now to internal waves, we can distinguishbetween two quite different effects on vertical profiles:(1) an (irreversible) microstructure and fine structureassociated with intermittent internal wave breaking,and (2) a (reversible) fine structure due to the verticalstraining of an otherwise smooth profile by internalwaves of short vertical wavelength. The reversible con-tribution to fine structure by internal waves was firstnoticed by Lazier (1973b) and Garrett (1973). How arewe to distinguish it from diffusive land other irrever-sible) fine structure?Let 8T and 8S designate departures in (potential) tem-perature and salinity from some long-time or long-dis-tance averages To(z), S0(z) at the same depth. Then$p = -a 8T + b Sis the associated density departure, with a(T, S.p) andb(T, S, p) designating the (positive) coefficients of ther-

    28IInternal Waves

  • 8/13/2019 Munk1981 Wunsch Chapter9

    19/28

    mal expansion and haline contraction. Take first thecase of an intrusion only (figure 9.19). If it is totallycompensated,8p =0,and if it is not totally compensated, it soon will be (ina time of order N-'). It follows that any vertical dis-placement of the isopycnals is not intrusive but due toa vertical displacement of the water, which we asso-ciate with internal waves (see next section). The ver-tical displacement 4 can be found from conservation ofpotential density:P(z + ) = PolZ).

    For the case of internal waves only, conservation ofpotential temperature and salinity giveT(z + ) = To(z), S(z + ) = So(z),and the -values from the three preceding equationsshould be the same:P = T = S- (9.17)

    Then in general, ,p ives the vertical displacement byinternal waves, and rT p and s - ,p re measures ofintrusive activity.Figure 9.20 shows the situation in (T, S)-space. In thecombined case, a projection parallel to the isopycnalscan separate the two effects. For constant a and b, it isconvenient to introduce a family of lines that are or-thogonal to the lines of equal potential density (Ve-ronis, 1972). They are here designated by r, for "spici-ness" (hot and salty ), and they give a measure of thestrength of the intrusion. The construction in (p, r)-space has some convenient properties. If the x- and y-

    internal wave intrusive glob

    T _7--

    h mp

    -- o,,--"-. ....... /

    P

    Figure 9.I9 Contours of potential temperature, salinity andpotential density in a vertical section (x,z) for an internalwave hump and a compensated warm and salty intrusive glob.

    axes are scaled in equivalent density units, bS and aT,then8p = -aT + bS, 87r= aT + bS. (9.18)

    Figure 9.21 shows plots of the inferred vertical dis-placements in an area 200 miles southwest of SanDiego (Johnson, Cox, and Gallagher, 1978). T, s, andUp hould all be alike for the case of a fine structuredue to internal waves only [equation (9.17)], and thisturns out to be the case down to a depth of 225 m.There is a broad intrusion between 225 and 260 m, anda narrow intrusion at 275 m. From a spectral analysisit was found that internal waves dominated the finestructure for all vertical scales that could be resolved,that is, down to 5 m.The displacement spectrum in vertical wavenumberm steepens from approximately m-2 for m < mu tom-3for m > m,,, with mu, near 0.6m-1 (-0.1 cpm). Thiskink appears to be a common feature in temperaturespectra (Gregg, 1977; Hayes, 1978), and is most clearlyportrayed in the temperature-gradient spectra figure9.22). A similar steepening is found in the spectrum ofcurrents and current shear, but at a somewhat lowervertical wavenumber (Hogg, Katz, and Sanford, 1978).A free-fall instrument called the "camel" for meas-uring the velocity microstructure has been developedby Osborn (1974; see chapter 14). Figure 9.23 presentsmeasurements in the Atlantic Equatorial Undercurrentduring the GATE experiment (Crawford and Osbom,1980). The most intense microstructure of temperatureand current was found above the velocity core. Themicrostructure in the core was weak and intermittent.Moderately intensive microstructure was found belowthe core, near the base of the thermocline. This isshown in great detail in figure 9.24. As an example,between 81 and 82 m there is an active temperaturemicrostructure with positive and negative OzT,accom-panied by an active velocity microstructure. Similarevidence is found in horizontal tows, as for example inthe upper part of figure 9.25 (Gibson, Schedvin, andWashburn, personal communication; see Gibson,1980). The important conclusion is that velocity mi-crostructure and small-scale temperature inversionsmust be closely linked, for one is not found withoutthe other.Occasionally one encounters patches of temperaturemicrostructure without velocity microstructure. Theinference is that these patches are the remains of amixing event for which the velocity microstructurehas decayed (fossil turbulence). Examples are found inthe vertical profiles (figure 9.24 between 69 and 70 mdepth), and in the horizontal tows (figure 9.25, bottom).But for the vertical profiles the temperature micro-structure is here limited to only positive OzT; the au-thors suggest that this might be a peculiarity of thecore (velocity and salt) of the undercurrent.282Walter Munk

    -----

    -

    i _

  • 8/13/2019 Munk1981 Wunsch Chapter9

    20/28

    internal wave

    T

    intrusion internal waveplus intrusion

    S SFigure 9.20 T-S relations for an internal wave hump and acompensated intrusive glob. The dots (e) correspond to theundisturbed positions of the five contours in figure 9.19. The

    open circles (o) give the positions through the center of thedisturbance. The "isospiceness" lines (constant r)are orthog-onal to the isopycnals (constant p).

    - T-30 0 30

    100

    U)

    -o-,"(200

    30C

    I

    3p-30 0 30I I

    Es- p-30 0 30I I I

    F-30 0 30

    Es

    4

    il-30 0 30CT 5p

    Figure 9.21 Displacement profiles in meters as inferred fromthe temperature, salinity, and density profiles. These shouldbe alike for internal wave produced fine structure. The meanT and S gradients were of opposite sign as in figure 9.19),hence the opposite signs of AT - ,p nd ts - ,. (Johnson, Cox,and Gallagher, 1978.)

    (C m-1)2c0-5pm

    in-B 10 10' 101m (cpm)Figure 9.22 Temperature gradient spectra for two stations inthe North Pacific. (Gregg, 1977.)

    Crawford and Osborn have calculated the dissipationrates. Typical values just beneath the core of the un-dercurrent range from e = 10- 4 to 10-3 cm2 s-3 (10-5 to10-4 W m-3 ). (Measurements away from the equator fallwithin the same limits.) But Belyaev, Lubimtzev, andOzmidov (1975) obtain dissipation rates in the area ofthe undercurrent from horizontal tows that are higherby two orders of magnitude."1The question of the relative magnitudes of verticaland horizontal scales has been examined by Hacker(1973) from a comparison of wing tip and nose tem-peratures of a rotating free-fall instrument. The hori-zontal separation is 1.7 m. The two records are coher-ent for vertical wavelengths down to 1 m. At smallerwavelengths the analysis is made difficult by the ran-dom tilts (50 ms) associated with internal waves. Byselectively analyzing depth ranges of small tilts, Elliottand Oakey (1975) found coherence over a horizontalspacing of 0.5-m down to 10-cm vertical wavelengths.The conclusion is that anisotropy extends beyond thefine structure into the microstructure, perhaps as faras the dissipation scale (-1 cm).The picture that emerges is one of a fine structurethat is usually dominated by internal wave strainingand is fairly uniform, in contrast to a microstructurethat is extremely patchy and variable even in the mean.Patches of temperature microstructure without veloc-ity microstructure ("fossil turbulence") evidently markthe demise of internal waves that had previously bro-ken.9.9 An Inconclusive DiscussionIs there a connection between internal wave activity,dissipation, and buoyancy flux? What is the explana-tion for the seeming steadiness of the internal wavefield? Having gone this far, I cannot refrain from con-tinuing with some speculation. The reader is encour-aged to go no further (if he has gotten this far).283Internal Waves

    S

    -~ ---

    A Tasadoy 1o Southtow13I , ,, ,,I

    "I.~q .. . .,,, ........... , V V * TT1,, .,,., ..... .......

    n

    I

    o-3L

  • 8/13/2019 Munk1981 Wunsch Chapter9

    21/28

    temp.(IC)10 20

    f/4. 924

    E-W current dzT(cm s - l) (C/cm)0 60 -0.1 0 0.1

    NA

    -1

    azU(s-I)1 0 11 1

    z U2(s-I)-1 0 1IT

    .1

    L

    Figure 9.23 Temperature and velocity microstructure in theAtlantic Equatorial Undercurrent at 0 18'S, 2801'W. (Craw-ford and Osborn, 1980.) The large scale-current profile wasmeasured by J. Bruce. The region between 65 and 82m isshown on an enlarged scale in figure 9.24.

    0

    i /t1

    200meters

    400 600i ].17ic a ti.c]temp. _ctivetm

    current ] 4 cm/s

    0 200 400I I econdssecondstemp.

    currentLI.... . . .1 ....

    0

    It' L tI

    ] 4 cm/s , .

    100 200meters 300 400

    Figure 9.25 Active and fossil turbulence from towed bodymeasurements during the mixed-layer experiment (MILE) nSeptember 1977 near ocean station PAPA. The body wastowed in the seasonal thermocline at a depth of 33 m; notethe horizontal temperature change by about 0.3C across thepatches. The 1C cale (left)refers to frequencies f < 1 Hz; the0.17C scale is for 1-12 H.(I am indebted to C. Gibson, J. Schedvin, and L. Washburn forpermission to show these measurements. See also Gibson,1980.)

    Figure 9.24 An enlarged section of the microstructure profileshown in figure 9.23. The encircled features have been attrib-uted to various sources of instrumental noise.

    284Walter Munk

    u

    -Z

    50E-C,.a)n

    100

    150I

    II

    dzT(C/cm)-.05 0 .05

    az U2(s-I)-0.1 0.1I I i, I65-Z

    I 75

    E,--a.a)C

    70

    fossil...d

    80

    I I-0o. o 0.1(SI )

    dzui

    .itI

    L.

    __

    _

    .&- --- .~ .M -itL OONOm, 4 OF ji NM'! v- L 10 M ML i A , A , * . N O. Ml I* M.I I I I I I I I I~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~

    Ij '

    n_

    Ir

    _

    j_

    7 L

    , ,

    r---

    I ].17cTMllI

    -i

    - I

    I

    I

    1_

    p

    IP

    r

    nonise

  • 8/13/2019 Munk1981 Wunsch Chapter9

    22/28

    We shall need some quantitative guidance for inter-nal wave intensities. For that purpose I shall use themodel spectrum of Garrett and Munk (1972b, 1975,1979),with slight modifications. The model fails nearthe ocean boundaries12 (Pinkel, 1975). The spectrumwas developed on the basis of rank empiricism, withno trace of underlying theory. But it has since gainedsome respectability by the theoretical findings of Wat-son and collaborators that the shape of the GM spec-trum is stable to nonlinear interactions, except for thelowest modes and near-inertial frequencies [Meiss,Pomphrey, and Watson (1979);Pomphrey, Meiss, andWatson (1980); see also McComas (1977)].9.9.1 Model Spectrum GM79The internal wave energy is assumed to be equallydistributed in all horizontal directions, so that only asingle horizontal wavenumber, k = (k2 + k2) 2, is used.Upward and downward energy flux are taken as equal.The spectra of vertical displacement, horizontal veloc-ity, and energy per unit mass are 3F(wo,j) = b

    2NoN-(o2 - f2)t- 2E({,j), (9.19)F(o,j) = FU + F.2 = b2NoN(Wo2 f2)c- 2E(w,j), (9.20)Fe(to,j)= (FU+ N2F) = b2NoNE(o,j), (9.21)where j is the vertical mode number, b 1.3 km thee-folding scale of N(z), with No - 5.2 x 10-3s - 1 (3 cph)the surface-extrapolated buoyancy frequency and f =7.3 x 10-5 s -I the Coriolis frequency at lat. 30 . We canignore Fwcompared to Fu. At high frequencies, o > f,kinetic and potential energy densities are equal: IFu =iN2 Fc. E(co,j) s a dimensionless energy density that isfactored as follows:Elo,j) = B(o)-H(j)-E,B(w) = 2 1r-fo-'l( 2 - f2)-12,

    (j2 + j2 )-H(j) = (j2 + )-'fB(o)d = 1,

    H(j) = 1.j=1The factor (W - f2)-112 in the expression for B(w) is acrude attempt to allow for the peak at the inertialturning frequency (see figure 9.11); ji = 3 is a modescale number, and E is the internal wave "energy pa-rameter." We setE = 6.3 x 10- 5 (dimensionless). (9.22)There is a surprising universality 14 to the value of E(mostly within a factor of two).The transfer into (w,k)- or (o,m)-space is accom-plished by setting F(co,j) j = F(lok)dk = F(co,m) m,withm = k ( ) 1/2= -rb(N ) /2 (9.23a)

    for a slowly varying N(z), in accord with the WKBapproximation [equations (9.4) and (9.7)].For most pur-poses we can ignore the situation near the buoyancyturning frequency,15so thatm kN(o 2 - f2)-'1 - b-'(NIN . (9.23b)

    For the sake of simplicity, the energy spectrum hasbeen factored into B(w)-H(j). But there is evidence fromPinkel (1975) and from the IWEX measurements(Miller, Olbers, and Willebrand, 1978) that there isrelatively more energy in the low modes at high fre-quency, and this could account for the astounding ver-tical coherences found by Pinkel in the upper 400 m athigh frequencies.I have no doubt that further discrepancies will befound; still, I believe that the model can now giveuseful quantitative estimates. For example, accordingto "Fofonoff's rule" (he disclaims ownership), themean-square current within a 1-cph band centered at1 cph is 1 cm2s-2; this compares to F(1 cph) =0.8 cm2 s-2 (cph)-l from (9.20). This agreement is notan accident, of course, the GM model having beenbased, in part, on the site D measurements (Fofonoff,1969).The mean-square quantitites are likewise in accordwith the usual experience. From (9.19), (9.20), and(9.21),

    = f d F = 3(Nw mj)= b2ENoN-' = 53(NINO)-l M2 ,

    (u2) = (u2) + (U2) = f do F(,i)= {b2ENoN = 44(N/No)cm2 s-2,

    ez) = dYo Fe()= b 2ENoN = 30(NINo) cm 2 s-2,

    (9.24)

    giving 7 m for the rms vertical displacement and7 cms - I for the rms current in the upper oceans be-neath the mixed layer. The energy can be written al-ternatively E(z) = [(u2) + N2(~)] so that the totalkinetic energy is three times the total potential energyin the GM model. pP(z) is the energy per unit volume;the energy per unit area isp = f pE(z)dz =pb2ENo Ndz = pb2ENo bdN

    - pb3NIE= 3.8 x 106ergcm- 2 = 3800 Jm -2 ,

    using b - = N-1 dN/dz as definition for the e-foldingscale b.9.9.2 UniversalityIt has turned out, quite unexpectedly, that the intens-ities are remarkably uniform in space and time.

    285Internal Waves

  • 8/13/2019 Munk1981 Wunsch Chapter9

    23/28

    Wunsch (1976) has made a deliberate attempt to findsystematic deviations for a variety of deep water loca-tions in the North Atlantic, with the purpose of iden-tifying sources and sinks of internal wave energy. Us-ing the frequency band to cph as a standard, theonly clear deviations he could find were associatedwith topographic features, particularly Muir Seamount,and even these were inconspicuous at short distance.In a further study (Wunsch and Webb, 1979) some evi-dence is presented for deviations on the equator and inregions of high mean shear.

    Figure 9.26 shows a continuing spectral display overan 18-day period, and this is found consistent with astationary Gaussian process (Cairns and Williams,1976). [The mean distribution is in accord with theequation (9.19) for Fco) = XFd(o,j),etting XH(j) = 1.]These observations were taken during 21 days of mildto moderate winds. Davis (personal communication)has recorded currents in the seasonal thermocline overa 19-day interval with two periods of heavy winds (fig-ure 9.27). The first event is followed in about 2 daysby an increase in mean-square currents, the secondevent in somewhat less time. Energy enhancement isby a factor of three or less. Johnson, Cbx, and Gallagher(1978) found a temporarily elevated spectral level on awindy day. Following these events, the intensities rap-idly relax to their normal state.9.9.3 GenerationThe observed growth times are consistent with a the-ory for the generation of internal waves by resonantinteraction with surface waves (Brekhovskikh, Gon-charov Kurtepov, and Naugol'nykh, 1972; Watson,West, and Cohen, 1976).But there are other means of generating internalwaves. Garrett (1979) has reviewed a variety of con-tenders, all of which fall (surprisingly) into the rightorder of magnitude. For reference, he takes 7 x10-3Wm -2 for the internal wave dissipation (corre-sponding to a relaxation time of one week). Globally,this amounts to 2 TW (terawatts: tera = 1012'.The totalloss of energy of the earth-moon system is known fromthe moon's orbit to be 4 TW, mostly by tidal dissipa-tion in the oceans. It is not impossible that surfacetides pump significant amounts of energy into internalwave motions via internal tides (cf. chapter 10). Othercontenders are surface forcing by traveling fluctuationsof wind stress and buoyancy flux, currents over bottomtopography, and extraction from the mean currentshear. There is no problem with supplying internalwaves with 2 TW of power; the problem is rather toeliminate some of the potential donors.9.9.4 InstabilityWe can now derive some numerical estimates for avariety of instability parameters. The spectrum Fu,,, of

    spectrum of vertical displacement

    ,, windstress' ,I7 19 , 251974 July

    Figure 9.26 Time-frequency display of F,(w) from MISERY 1and MISERY3. (Caims and Williams, 1976.) (t) is the depthof the 6.60" isotherm (at a mean depth of 350 m) in a location800 km offshore of San Diego, California, measured with ayo-yoing midwater capsule. The squared wind (bottom) showslight winds at the start and end of the experiments.

    32

    C,E

    30- kinetic energy2'~0- ~ 0.3-1.0 cph

    0GM

    v .- .. .... .. . , . . . . . I . . . . . .0-

    10-

    ,Ev

    42m

    GM- As L u N/V=_ 54 mkinetic energy1.0-2.5 cph

    G M- 1>l\,l U/__2-

    N 40

    2E

    GM- *

    42m

    54m

    0- wind stressO' A_

    . . . . ' 311 ' ' 'August 1977 SeptemberFigure 9.27 Kinetic energy (u 2) = (u) + (u2) during themixed-layer experiment (MILE) on station PAPA (50"N,145"W)at 42 and 54 m depths. The upper two plots refer toa frequency band of 0.3 to 1.0 cph, the next two plots to aband 1.0 to 2.5 cph. The squared wind (bottom) shows twoepisodes of large wind stress. The GM model levels are indi-cated [using N = 0.023, 0.0096 s-1 (13, 5.5 cph) at 42, 54 m],though the model is not really applicable to such shallowdepths and sharp N-gradients. (I am indebted to R. Davis forpermission to use his measurements.)

    286Walter Munk

    .

    June

    1

    1

    r ..A

    2' 8

  • 8/13/2019 Munk1981 Wunsch Chapter9

    24/28

    the reciprocal Richardson number is defined byRi - = ((u')2)/N2 = dco Fu'/N(O,j),where ((u')2) = ((dzu,)2) + ((Zu2)2) = m2(u2). Butm = 7rrb-'(N/No)j (9.23b) and soFu,/N= 2*rE(N/No)f(o2 f2)o-3(o2 - f2)-1/2j2H(j). (9.25)The principal contribution comes from the inertial fre-quencies. Performing the co-integration,Ri-' = 'Tr2E(N/No)j2H(j). (9.26)We now perform the mode summations (subscript ufor upper)juY j2(j2 + ,)-' - Ju,j=ljUE (j2 j2)-i lij2 rj* - 1) -1' = 0.47j=1for ju >> ij = 3. The spectrum Fu,/N(j)is white (exceptfor the lowest few mode numbers) and Ri-' depends onthe choice of the upper cutoff ju. In terms of the lim-iting vertical wavenumber mu = 7rb-l(N/N)ju we have,finally,Ri - ' = urJEmub. (9.27)If we identify mu with the kink (figure 9.22) at 0.6m-'(0.1 cpm), then Ri-' = 0.52. If the spectrum is extendedbeyond the break, with a slope m - 1 to some new upperlimit muu = 10mu (say), then Ri - = 0.52(1 + lnmuu/m,)= 1.72. Hogg, Katz, and Sanford (1978) find Ri-' near0.5 in the open ocean near Bermuda, including m up to0.2 cpm. An interesting scatter plot has been producedby Eriksen (1978) and is shown in figure 9.28. Theconclusion is that the internal-wave shear field is as-sociated with Richardson numbers of order 1:Ri-' = order(l). (9.28)

    We can proceed in a similar manner with regard toadvective instability. The simplest generalization ofthe earlier discussion (section 9.7) is to derive the spec-tral decomposition of (u2/c2), c = co/k. (Would(u2)/(C2) be better?) The equations of continuity andof dispersion (away from the turning frequencies) canbe writtenkf+m(4=0, c = ol/k=N/m,so thatu mu u'c mc N'Similarly, for horizontal and vertical strain,a/1ax = -/az = ik6 = (k/o)u = u/c.

    0.510- -

    dN 2

    0 0.5x10 -

  • 8/13/2019 Munk1981 Wunsch Chapter9

    25/28

    From the expressions (9.24) for (2) and (u2), eithercondition (9.30) or condition (9.31) leads to an energyparameterE = [(02)ICC']2No/N. (9.32)For C = 5, C' = 6, E = 10-3(4)2)2No/N. Then with (42)a moderate fraction of its critical value 1, we can re-cover the numerical value E = 6 x 10-5. This is not tosay that E has been calculated from first principles; itis only to say that acceptable values for the variouscoefficients lead to a small numerical value of the di-mensionless energy parameter, as observed.9.9.6 DissipationFor small numerical values of the instability parameter(42) we are in a regime of sparse instabilities in spaceand time (such as incipient whitecaps in light winds).When (4)2) is near 1 the probability for instabilities ishigh. For an a priori estimate of (42) we require (1) amodel to relate (4)2) to the internal wave energy dis-sipation, and (2) an estimate of the rate of dissipation(or generation for a given steady state). This is essen-tially the procedure followed by Longuet-Higgins(1969a) n his stimulating attempt to interpret the Phil-lips saturation constant for surface waves.Perhaps the simplest scheme is to relate the dissi-pation to the probability for 4) > 1.16 The variance of4 associated with 4)> 1 is (2) p ( > 1) for uncorre-lated and 4 (as when 4 = 0z,). Potential energy isproportional to (2); accordingly the rate of fractionalenergy dissipation can be written1 dE = -op() > 1), (9.33)E dtwhere -' is the characteristic interval during whichthe energy associated with 4) > 1 is lost to the organizedwave field and renewed by generation processes. For arough estimate (Garrett and Munk, 1972a),

    N N,2 _,-2tf/F,(&o)do,/ Fl(wo)do

    7 -2fN (9.34)for any of the 4-spectra [such as (9.25)]. A Gaussian 4)-distribution p(4)) = r-1/23exp(-i,34) leads to

    1p (4) > 1) = r-' 2/3-1/2xp(-3), 2 ' (9.35)provided ,3 is large. In the upper ocean or= 20 per day,and 3 = 2.1, 3.7, for relaxation times of 1 day, 1 week,respectively. The foregoing numerical values are notimportant; what is significant is that a tenfold increasein the rate of dissipation (and generation) is accom-panied by only a threefold increase in /32 (and hence inwave energy). Thus the energy level stays within rathernarrow limits even though generation and dissipation

    processes may vary widely, particularly for large /3. Wepropose for a "universality hypothesis" that the energylevel responds only logarithmically to variable forcing.We shall examine this situation in more detail.9.9.7 The Energy BalanceThe differential equation of wave energy can be writtendEldt = G(t) - D(t),where G(t) and D(t) are the rates of energy generationand dissipation. We use the notation E, , D to repre-sent the "normal" state of internal wave statistics.From (9.32) and (9.35)E - 3-2(No/IN), D - fi-'/2e-,with ID = . We define the relaxation time

    = G ;accordingly t-' is the initial rate of decay for a wavefield in equilibrium with 0 if the generation is sud-denly turned off. The differential energy equation cannow be written+4 xp[(l --12)] =g(r (9.36)h 2(N/No) - '2,where(T) = E/E, g(-) = G/G,with = t,(g) = 1, (g) = 1.For large , the dissipation is g14 exp and thus large;for small , it is Wg,4exp _(W-12) and thus very small.The problem is to derive properties of the energystatistics for given generation statistics. (This is relatedto the fluctuation-dissipation theorem in the study ofBrownian motion.) Two special solutions are easilyfound. For an equilibrium situation, (9.36) withdWl dr = 0 gives the values g(g;i) in table 9.2. Depar-tures from the normal state in E are much smaller thanthose in g, particularly at large 3 and for small g's.To obtain some feeling for the nonlinear responsetime, let g(r) go abruptly from 1 to g at time 0, and set

    (r) = 1 + E(T),with E

  • 8/13/2019 Munk1981 Wunsch Chapter9

    26/28

    Table 9.2 Energy Equilibrium i({g,) from (9.36) for dI/dT =0 (top), Dimensionless Response Time r1/2 (9.38) for anAbrupt Change in Generation from 1 to g (center),Relaxation Time r,12 9.39) for a Change fromg to 1(bottom)a

    g (dB)-10

    Energy levels in dE2 -6.0

    A8 5 -3.010 -1.7

    -5 0 5 10

    -3.4 0 4.9-1.7 0 2.0-0.9 0 1.0

    Response time (dimensionless)2 2.75 1.41 0.69

    A 5 1.10 0.56 0.2810 0.55 0.28 0.14Relaxation time (dimensionless)

    2 1.43 1.02 0.69A 5 0.57 0.41 0.28

    10 0.29 0.20 0.14

    1 = (r1)- m12 exp( -,8)A,where p1(z), (z) are the dissipation and energy perunit volume, respectively. The dependence on depth isthrough(r-n, n- l'2, A n,with n(z) = N(z)/No; hence

    13.0 /1 = od-3912xp(,/0 - ), A = 0oP-2.4.62.1

    0.32 0.140.13 0.060.06 0.03

    0.45 0.270.18 0.110.09 0.05

    a. Generation g and energy g are in decibels relative tonormal levels.

    in accordwith d/edr = -1 for the initial rate of decay.As T goes from 0 to o, z goes from 1 to g, and e from0 to 2 - 1ng. Half energy response is for E = -1 lng,z = g1 2, andr1/2= 2-lg -1 n(l + g 2). (9.38)For energy relaxation, going abruptly from g to 1 attime 0, the solution is

    T =ln, - 1)z 1/2= 2-'ln(1 +g- 2). (9.39)= g(z - 1)'Half-times T1/2are given in table 9.2. For a linearsystem these would all be the same. Here the times areshorter for internal wave storms (g > 1) and longer forcalms (g < 1), and this variation is more pronouncedin response to a change in generation from a normal toa perturbed level than for relaxation back to normalgeneration. (The assumption of weak nonlinearity forcomputing T is violated in the columns forg = +10 dB.)The imperceptible decay of internal wave intensitiesduring relatively low winds (figure 9.26) is consistentwith half the normal energy, and the rapid decay fol-lowing a blow (figure 9.27) requires perhaps three timesnormal energy. If we can count on a storm or someother generation event to "top up" the internal waveenergy once every hundred days, then we may expectthe wave energies to remain generally within a factorof two.I have paid no attention to depth dependence. Thedissipation can be written

    Writing dz = b dnln = -2b d/3l/,f D dz = 2bof 11/2 exp1,3-l8) dl3,f dz = 2bAof -3 dl.

    Integrating from the surface ( = o)(3 oo), he integrals for large o (assumed) givef15z = 2bO,lo30-2,to the bottompreviously as-

    f dz = boi 2with an "integral relaxation time"1l721olbo jgl:/ 04eool-. 19.40)

    Suppose the generation takes place in the upper fewhundred meters, so that the dissipation in the interiorocean is compensated by downward radiation of inter-nal wave energy. From a rotary decomposition of cur-rent profiles, Leaman (1976) estimates a downward fluxof order 10- 4 Wm-2 . We compare this with the inte-grated dissipation beneath a scale depth z = -b --1.3 km where n = e-Land 8 = 8oVe):P B l(z)dz = 2 -11e-ll 4 a-6e-ooN20pb3a,where P is the "normal" dimensionless energy param-eter. The result is 130 2.1. The corresponding integralrelaxation time [equation (9.40)]is 7 days. The normalsurface relaxation time isio = A 0o = (Tr3plo12ela-' = 1.1 days.The dimensionless times in table 9.2 can be interpretedas shallow response times in days.The strong dependence of dissipation on depth is aninherent feature of the proposed phenomenology.9.9.8 MixingThe balance between production and dissipation of tur-bulent energy can be represented bytX = E = E + Ek- (9.41)pep is the rate of production of potential energy, e.g.,the buoyancyflux g(w'p'), with the primes designat-ing the fluctuating components; pek is the dissipationof kinetic energy into heat. The fraction of work going

    289Internal Waves

  • 8/13/2019 Munk1981 Wunsch Chapter9

    27/28

    into potential energy is the flux Richardson numberRf:Rf = Ep/p = E,/E. (9.42)Rf < 1 in order for Ek> 0. We have previously discussedthe evidence that something less than of the kineticenergy dissipated appears as potential energy. Takinga typical value Ep/Ek = (Thorpe, 1973b, p. 749) givesRf = . The eddy flux of density can be written in termsof the eddy diffusivity A: (w'p') = A dp/dz =AN2 (g/p)-l = pp/g, and so

    EsE Ek RfA = - N- R f N2f (9.43)Our procedure is to estimate E, from internal wavebreaking, and to compute A and E from (9.43), usingRf = . There are, of course, other sources of turbu-

    lence; Osborn (1980) stresses the work done by theambient turbulence in a mean shear: tz = (u'w') du/dz.His procedure is to estimate from the measuredmean-square shear, and to compute A from (9.43).The random superposition of internal waves leads tothe intermittent occurrenceof "traumata" 7 associatedwith > 1. The traumata are the locus of incipientturbulence, and quickly spread to some thickness Awithin which the average , is reduced from 1 to about0.9 (Thorpe, 1973b). Subsequently the patch continuesto grow to some maximum thickness Amaxat the timethe surrounding is largest, always keeping withinthe patch to 0.9.

    The change of potential energy per unit surfacearea associated with perfect mixing over a depth Ain a density gradient dp/dz is (1/12)g(dpldz)A =(1/12)pN A . (Imperfect mixing just reduces the factor1/12.) The average change of potential energy per unittime per unit volume is thenPE, = 1 pN2 A3 v,12where v is the number of traumata per unit (t,z)-space.pe, equals the buoyancy flux pN2A by definition of theeddy diffusivity A; henceA 1 A3 v12I have previously identified A with mu' [equation(9.31)]. For v we writeV= m'op( > 1), m' = 0.2m, 7r-'(fN)where m' is the rms spacial frequency derived frcequation analogous to (9.34). Putting all this togA 10-2 A2/T, T o-lflm'12e,

    1/2

    )m an

    where T is the expected time interval between eventsover a distance I/m'. This is of similar form as theresult of Stommel and Fedorov (1967),as is inevitablefor what is, after all, a mixing-length theory.The principal conclusion is a strong dependence ofA on depth, and on any departures from normal gen-eration. The numerical value for the normal state isA = 10-2 cm 2 s -', much lower than the global 1 cm 2 s- '.9.9.9 Saturation SpectraThere is an essential distinction between the usualformulation of turbulence and the saturation processesas here envisioned. We consider the regime of sparseinstabilities in space and time (such as incipient white-caps in light winds). Then the +-field consists of scat-tered and uncorrelated spikes, and the -spectrum isaccordingly white up to some limit that is the recip-rocal of the vertical extent of the spikes. The dissipa-tion is localized in physical space, and thereforebroadly distributed in wavenumber space. In the usualturbulent situation, the dissipation is confined to anarrow (dissipation) region in wavenumber space, andspread in physical space.A white spectrum in any of the +-spectra, whethershear, advection or strain, implies an m- 2 energy spec-trum. The energy spectrum steepens (perhaps to m -3 )in the transition from the intrinsic to the compliantwaves. Presumably the m - 3 energy spectrum extendsto the Ozmidov (or Richardson or Monin-Obukov)scale m, = (N3/e)'2 4 m 1 (about 0.6 cpm), which isconveniently close to the definition of the microstruc-ture boundary, then flattens out to the Kolmogorovdissipation scale mk = (E/v3)'4 3 cm ' (0.5 cpcm), andfinally cuts off exponentially (Gregg,Cox, and Hacker,1973, figure 11). But such a description of "in-the-mean" scales may not be appropriate to a patchy en-vironment, and is anyway beyond the scope of thissurvey.9.10 ConclusionI shall end as I started: the connection between internalwaves and small scale processes-that is where the keyis. I feel that we are close to having these pieces fallinto place, and I am uncomfortable with having at-tempted a survey at this time.Notes1. They were found at Loch Ness at about the same time(Watson, 1904; Wedderbum, 1907).

    ether, 2. The temperature measurements were made using a sub-marine cable from a recording Wheatstone bridge on the shore(9.44) at Bermuda to two resistance thermometers offshore: both layon the bottom, one at a depth of 50 m and the other at 500 m(Haurwitz, Stommel, and Munk, 1959).

    290Walter Munk

  • 8/13/2019 Munk1981 Wunsch Chapter9

    28/28

    3. Salinity is not directly measured, but has to be inferredfrom conductivity or sound speed, which are primarily re-sponsive to temperature. This large temperature "correction"has been a source of some difficulty.4. The notation U' = 8zU refers to the ambient shear, andu' = 0,u to the shear induced by the orbital wave motion. Thedistinction is not always so clear.5. In general, forward flow refers to positive c(g x curlU),where g and U are the vectors of gravity and ambient velocity,respectively.6. I am indebted to D. Broutman for very considerable im-provements of this section, and for the preparation of figure9.14.7. The linearized calculations indicated that they are in factnot small.8. Backward breaking occurs for negative shear; this can bevisualized by turning the figure upside down.9. Frankignoul's (1972) treatment of fnite amplitude waves[his equations (24)and (28)] ead to precisely the same result.10. Garrett points out that a lot of laboratory experimentshave been sweet-and-sour rather than spicy.11. It has been suggested that vibration and temperature con-tamination contributes to the high values from the toweddevices; it has also been suggested that the dropped deviceshave inadequate dynamic range to measure in the highlyactive patches where most of the dissipation takes place.12. A normal-mode formulation is applicable near the bound-aries (Watson, Siegmann, and Jacobson, 1977).13. Frequencies are in rads-', wavenumbers in radm-1 . Forcomparison with computed spectra we sometimes include (inparenthesis) the values in cycles per hour (cph) and cycles permeter (cpm).14. We note that Fe(t >> f) a E(w,j) - o-2fE. There is someevidence that the spectral energy density is independent oflatitude (Wunsch and Webb, 1979; Eriksen, 1980), and weshould probably replace fE by NoE', with E' = (f30oNo)E =8.8 x 10- 7 the appropriate f-scaled energy parameter.15. Desaubies (1973, 1975)explains the observed N-peak inthe spectrum and the vertical coherence of vertical displace-ment.16. This is related to the "intermittency index" evaluated byThorpe (1977)from temperature inversions in Loch Ness.17. "A disorderly state resulting from stress." This descrip-tive terminology is due to McEwan (1973).