943 NAVAL POSTGRADUATE SCHOOL Monterey, California THESIS OBSERVATIONS OF INERTIO- GRAVITY WAVES WAKE OF HURRICANE FREDERIC IN THE by Lynn K. Shay December 1983 Thesis Advisor: R. L. El sb( ?rry Approved for public release; distribution unlimited T215693 Unclassified SECURITY CLASSIFICATION OF THIS PAGE C*hon Data Entered) REPORT DOCUMENTATION PAGE T~REPO»T NUMBER READ INSTRUCTIONS BEFORE COMPLETING FORM 2. GOVT ACCESSION NO 3. RECIPIENT'S CATALOG NUMBER 4. TITLE (and Subtitle) Observations of Iner tio-Gravity Waves in the Wake of Hurricane Frederic 5. TYPE OF REPORT & PERIOD COVERED Master ' s Thesis ; December 19 8 3 6. PERFORMING ORG. REPORT NUMBER 7. AUTHOR^ Lynn Keith Shay 8. CONTRACT OR GRANT NUMBERf*.* t. PERFORMING ORGANIZATION NAME ANO ADDRESS Naval Postgraduate School Monterey, CA 93943 10. PROGRAM ELEMENT. PROJECT, TASK AREA 4 WORK UNIT NUMBERS II. CONTROLLING OFFICE NAME ANO AOORESS Naval Postgraduate School Monterey, CA 93943 12. REPORT DATE December 1983 13. NUM8E. iff 5 PA GES 14. MONITORING AGENCY NAME ft AOORESSf'f dIUatant /ram Controlling Ollice) 15. SECURITY CLASS. (o( ihla report) Unclassified I5«. DECLASSIFICATION/ DOWNGRADING SCHEDULE IS. DISTRIBUTION STATEMENT (ol thla Report) Approved for public release, distribution unlimited 17. DISTRIBUTION STATEMENT (ol (ft* abetrect entered In Block 20, II different from Report) IS. SUPPLEMENTARY NOTES IS. KEY WOROS (Continue on tereree mid* II neceeamty and Idontlty by block number) hurricane Frederic, iner tio-gravity waves, inertial waves, group velocity, phase velocity, normal modes, barotropic, baroclinic, complex demodulation, least squares, Brunt Vaisala frequency 20. ABSTRACT < Continue on reveree eide II nacaaaary and Identity by block number) Inertial waves excited in the mixed layer by hurricane Frederic, had horizontal scales of approximately 1 to 2 times the baroclinic Rossby radius of deformation (50 km) of the first mode near the DeSoto Canyon. Initially, energy propagated vertically at about 1.25 km/d and horizontally at about 80 km/d. These waves spun down over e-foldine scales of four inertial periods as energy propagated vertically at 270 m/d and horizontal- DO,: FORM AN 73 1473 EDITION OF 1 NOV «S IS OBSOLETE S/N 0102- LF- 014-6601 Unclassified SECURITY CLASSIFICATION OF THIS PAGE (When Data Entered SECURITY CLASSIFICATION OF THIS PAGE (TWi«n Dmtm Entfd} ly at 30 km/d. Inert io-gravity waves in the deep thermocline had horizontal scales of 25 to 50 km and vertical scales approximate- ly equal to the water depth. The energy of these waves was domin- ated by the barotropic mode with some contributions from modes 1 and 2. These waves were not admitted to the shelf region because the bottom slope was greater than the slope of the internal wave characteristics. The mean flow followed the isobaths at all levels, but it wa^ in the opposite direction in the bottom layer. The mean flow initially decreased along the eastern boundary of the canyon as the storm forcing readjusted the flow. Near-bo t torn temperature variations of 4*C were associated with the storm surge and advec- tion in the along-track direction, particularly along the north rim of the canyon. S-N 0102- LF- 014- 6601 Unclassified SECURITY CLASSIFICATION OF THIS P kG€(Wtf€i Dmtm Bnffd) 2 Approved for public release; distribution unlimited Cfcservaticns of Inert ic-Gravity Waves in the Wake of Hurricane Frederic by B. S. Oceanocrap Lync K. Shay hy, Florida Ins'titut e of Technology, 1976 Submitted in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE IN OCEANOGRAPHY from the NAVAL POSTGRADUATE SCHOOL December 1983 D'JI ABSTRACT .../. c ^ Inertial waves excited in the mixed layer by hurricane Frederic, had horizontal scales of approximately 1 to 2 times the tarcclinic Rossby radius of deformation (50 km) of the first mode near the CeSoto Canyon. Initially, energy propagated vertically at about 1.25 km/d and horizontally at 80 km/d. These waves spun down over e-folding scales of four inertial periods as energy propagated vertically at 270 m/d and horizontally at 30 km/d. Inertio-gravity waves in the deep therirccline had horizontal scales of 25 to 50 km and vertical scales approximately equal to the water depth. The energy of these waves was dominated by the barotropic mode with seme contributiens frcm modes 1 and 2. These waves were not admitted to the shelf region because the bot- tom slope was greater than the slope of the internal wave characteristics. The mean flew followed the isobaths at all levels, tut it was in the opposite direction in the bottom layer. The mean flow initially decreased alcng the eastern boundary of the canyon as the storm forcing readjusted the flow. Near- bottcm temperature variations of uoc were associated with the storm surge and advection in the along-track direction, particularly along the north rim of the canyon. TABLE OF CONTENTS I. INTRODUCTION 13 II. HISTORICAL REVIEW 16 III. DATA AND ANALYTICAL METHODS 23 A. CURFENT METER DATA 23 B. STCFM TRACK 24 C. WINE FIELD 27 D. VERTICAL TEMPERATURE GRADIENTS 27 E. TEMPORAL AND SPATIAL SCALES 30 F. CCMEUTATION OF SPECTRA 32 G. COMPLEX DEMODULATION 34 IV. STORM PEFICD 37 A. INERTIAL RESPONSE 37 B. SUBINERTIAL RESECNSE 63 C. SUFEFINERTIAL RESPONSE 73 V. NORMAL MODES 81 A. THECEY 81 B. EATA ANALYSIS 83 VI. CONCIUSICNS 95 APPENDIX A 98 A. CCMELEX DEMODULATION 98 B. POTENTIAL ENERGY 100 C. LEAST SQUARES FIT 101 APPENDIX E -10U A. DEEP REGION 104 B. SHAIIOW REGION 110 LIST OF REFERENCES 115 INITIAL DISTRIBUTION LIST 118 LIS! OF FIGURES Figure 1. Track of Hurricane Frederic 14 Figure 2. Satellite Photograph of Frederic at 2001 GMT 12 Sept. 1979 26 Figure 3. Wind Speed and Direction Measured by NOAA Eucy 42003 28 Figure 4. Vertical Temperature and Brunt-Vaisala Freguency Profiles 30 Figure 5. Freguency Response for the Band-pass and Lew-pass Filters 36 Figure 6. Band-Pass Filtered Mixed layer V Velocity Component at CMA3 36 Figure 7. Mixed Layer Velocity Components at CMA2. . . 38 Figure 8. Sinrilar to Fig. 7, except at CMA3 40 Figure 9. Time Series of HKE in the Mixed layer .... 41 Figure 10. Similar to Fig. 8f except in the deep ttermocline (251 m) 43 Figure 11. Progressive Vector Diagrams at CMA3 44 Figure 12. Time Series of HKE at CMA3 47 Figure 13. Normalized HKE Spectra from CMA3 49 Figure 14. Mixed Layer Normalized Rotary Spectra from CMA3 50 Figure 15. Vertical Scales at CMA3 55 Figure 16. Estimates of Hcrizontal Scales in the Mixed Layer 58 Figure 17. Hcrizontal Scales in the Deep Thermocline . . 60 Figure 18. Low-pass Filtered PV D at CMA2 66 Figure 19. Lew-pass Filtered PVD at CMA3 67 Figure 20. Thermocline Temperature Time Series from CMA3 at 251 m. 69 Figure 21. Temperature Tine Series from CMA2 at 324 m 71 Figure 22. Potential Energy Spectra from CMA3 72 Figure 23. Cress-Spectrum Between V-Velocity and Temperature 75 Figure 24. Mid-depth Velocity Components from CMA1. . . 76 Figure 25. Horizontal Velocity Eigenf unctions at 29.33 N and 87.11 9 84 Figure 26. V Velocity Coefficients at CMA3 86 Figure 27. Same as Fig. 26, except at CMA2 87 Figure 28. V Component Time Series at CMA3 93 Figure 29. Surface PVD fcr a) CMA2 and b) CMA3. ... 105 Figure 30. Bottom PVD for a) CMA2 and b) CMA3 106 Figure 31. Temperature Time Series from CMA3 at 457 m 111 Figure 32. Mid-depth PVC at CMA1 112 LIST OF TABLES TABLE I. Comparison of Inertial-Internal Wave Parameters 17 TABLE II. A Synopsis of Storm Period Observations . . 25 TABLE III. Rcssby Radii cf Deformation 32 TABLE IV. Normalized Rotary Spectrum Analysis at Inertial Frequency 51 TABLE V. Vertical and Horizontal Coherencies at the Inertial Frequency 53 TABLE VI. Comparison of Vertical and Horizontal Group Velocity 61 TABLE VII. Normalized Rotary Spectrum Analysis at Sufcinertial Frequency 70 TABLE VIII. Normalized HKE Spectra at Inertial and Semi-diurnal Frequency 77 TABLE IX. Normalized Rotary Spectra at the Semi- diurnal Frequency 79 TABLE X. Variance of the Normal Modes at CMA2 and CMA3 90 TABLE XI. Normalized HKE Spectral Estimates at 95% Confidence 107 TABLE XII. Normalized Rotary Spectrum Analysis . . . 109 TABLE XIII. Normalized HKE Spectral Estimates at 95% Confidence 113 ACKNOWLEDGEMENT The author extends his gratitude to all those who have contributed tc the evolution of this thesis. First and foremost, I would like to express my deepest appreciation to my faculty advisor, Erof. Russell Elsberry. His guidance and wisdom during a very difficult time of my life were exceeded only by his thorough knowledge of the subject mat- ter. Under his leadership, this thesis work has been the crescendo of my graduate education. The expertise of Dr. Andrew Willmott is held ir. high regard. His attention to mathematical details has irade this thesis a much better man- uscript. I also wish to thank Prof. Christopher Mooers for his willingness to share his knowledge on the application of normal mode theory to the analysis of current meter data, and fcr providing the computer program, IWEG, to solve for the vertical structure. His criticism and advice have con- tributad significantly to my development as an oceanogra- pher. The wisdom and guidance of Dr. William Hart whenever encouragement was needed has also been greatly appreciated. The dedicated efforts cf the personnel who maintain the computer systems at the W. R. Church Computer Center at NPS as well as at NAVOCEANO are greatly appreciated. Their wcrk certainly helped expedite the publishing of this thesis. 10 I also wish to acknowledge the unheralded efforts of the current meter group of NAVOCEANO . In particular, the fine work of Messrs. Jose Alonso and Steve Clark is appreciated and have made this thesis work a reality- The advice of Drs. Eill Wiseman (LSU) , Zack Ha llock (NORDA),and Ben Korgen (NAVOCEANO) and Mr. Jack Tamul (NAVOCEANO) was extremely helpful in the analysis cf the data. Drs. Hank Perkins (NORDA) and Ortwin vcn Zweck (NAVOCEANO) have shared some stimulating ideas concerning inertial-internal wave motions. I further extend my gratitude to Mr. Jerry Carroll for his assistance in gaining access to this very unique data set. The Physical Oceanography Branch at NAVOCEANO kindly pro- vided some of the computer software used in the data analy- sis. The expertise cf Mr. Steve Lauber in the pre-processing of the data greatly aided the data analysis. The drafting cf seme of the figures was done with great care by Ms. Viola Lutrell, Mr. Chuck Murphy, and Mr. Glenn Vocr- his cf NAVCCEAKC. I thank the two most important people in my life, my wife, Sally and daughter, Missy. Their patience and encour- agement during the evolution of the thesis has helped me a great deal. 11 Finally, I would like to thank NAVOCEANO for their sup- port ever the first three quarters cf my graduate studies Furthermore, the author extends his appreciation to ONR fo: support during the final stages of the research. 12 I. INTRODUCTION " the deer alone learnath " Nietzsche The oceanic response tc intense, transient atmospheric events, such as hurricanes, is dominated by robust insrtial wave excitation in the mixed layer and the subsequent propa- gation and dispersion of inertio-gravity waves in the ther- mocline. The study of these forced waves has been largely through numerical and analytical investigations, because of the lack of sufficient data to resolve the scales and ener- getics of the motion. Hence, observations of ocean currents and temperatures during a hurricane are neccessary for vali- dating models and theories dealing with forced inertio-grav- ity t>aves. A comprehensive data set of ocean observations was col- lected by the 0. S. Naval Oceanographic Office (NAVOCEANO) during the passage of hurricane Frederic in the summer of 1979 (Shay and Tamul, 1980). NAVOCEANO deployed three moored taut-wire current meter arrays in the northern Gulf of Mexico southeast of Mobile, Alabama. During the period of deployment, hurricane Frederic passed within 80 to 130 km 13 of the array sit€S (Fig. 1) at 2 100 GMT 12 September. Five hours later, Frederic made landfall at Dauphine Island, Ala- bama. During and subseguent to the passage of Frederic, the National Hurricane Research Division dropped several expend- able bathythermographs (AXET) from reconnassiance aircraft in the area of the current meter deployments (Black, 1983). SC.IE ft Figure 1. Track of Hurricane Frederic. The contours are in fathoms, A depicts positions of current meter arrays, (NAVOCEANO) , ■ depicts the position of AXET drops (Black, 1983) . The track of hurricane Frederic is based on a report by Hebert (1979) . 1U This thesis addresses the problem of the vertical and horizontal dispersion of the forced inertial wave energy from the mixed layer. The forcing readjusts the mean flow as well as generating anisotropic inertial motion. The cur- rent meter observations allow determination of the scales and the energetics of the forced, anisotropic wave motion through use cf spectrum analysis and complex demodulation. The modal structure is examined by formulating the Sturm- Liouville problem and thsn performing a least squares fit between the eigenf unctions and the demodulated time series of the current observations. 15 II. HISTORICAL REVIM A nonlinear model for a stationary hurricane (O'Brien and Reid, 1967) was developed to simulate the observations made in the wake of hurricane Hilda (Leipper, 1967). The results indicated that upwelling regions were restricted to the area enclosed by the iiaximum wind regime, whereas down- welling occurred outside this regime as warm water was advected away from the stcrm center. Cne of the most proncunced responses of the ocean to wind forcing is the generation of inertial oscillations in the surface mixed layer. Pollard (1970) simulated the gen- eration of these waves using a two-dimensional, wind-driven model developed by Pcllard and Millard (1970) . The pre- dicted amplitudes and decay rates cf the forced oscillations agreed well with the Woods Hole Oceanographic Institution (WHOI) * site D1 observations. Using the same data set. Pol- lard (1980) showed that under relatively strong wind condi- tions, 67 % of the horizontal kinetic energy (HKE) in the mixed layer was found near the inertial frequency. The sub- sequent downward propagation of energy was estimated to be 16 of the order cf 10~3 cm/s ( 1 m/d) . This vertical group velocity was too small to account for the radiation of ir.er- tial waves and the subsequent loss of energy from the mixed layer. The corresponding scales of the inerrial wave motion were about 100 to 240 d in the vertical and hundreds of kilometers in the horizontal (Table I). TABLE I Compariscn of Inert ial- Internal Wave Parameters I Length Brooks (1983) S cale s Horizontal (km) 370 Vertical (m) 1000 Group Velocity Horizontal (km/d) 23 Vertical (m/d) 60 From Brocks (1983) Price (1983) 480 1000 86 160 Pollard (1980) 700-1700 100-240 0. 8-17. 0.03-3. Dsing a linear, two-layer model, Geisler (1970) simu- lated inertic-gravity waves as part of the baroclinic response in the wake of a translating storm. The critical factors governing the generation of these waves were: • the translational speed of the hurricane must be greater than the internal wave phase speed; and • the horizontal scales of these waves must be comparable to the oceanic Rcssby radius of deformation. 17 He further noted that the ocean response to a moving hurri- cane is strcngly baroclinic. The ocean thermal response to hurricane forcing, was investigated numerically by imposing Ekman layer dynamics in a mixed layer model (Elsberry et al., 1976). The input of energy frcm the wind stress was forced to generate inertial oscillations and cause turbulent mixing via entrainment. Upwelling alsc enhances the turbulent mixing process in the upper layers of the ocean. The main result was that advec- tion dominated the thermal response near the storm track as opposed to the heat loss from the ocean surface to the storm. Strong atmospheric fcrcing enhances the turbulent mixing process in the surface mixed layer and causes it to deepen. As the mixed layer continues to deepen, the thermocline begins to erode as water from the thermocline is entrained into the mixed layer. However, stratification tends to sup- press turbulent mixing and aids in the creation of small- scale internal waves either through entrainment or Kelvin-Helmholtz instabilites (KH) at the base of the mixed layer (Pollard et al. , 1973 ; Garwood, 1977) . 18 Eelow the mixed layer, ocean current variability is linked to vertically pre Plating wave groups of the internal wave field (Kase and Gibers, 1979) . The actual mechanisms for the vertical transport of energy from the wind forced mixed layer tc the thermccline for the generation of large scale inertial-internal waves are not well understood. Per- haps the most effective way of creating these large scale inertial-internal waves is through Ekman suction. Krauss (1972 a,b, 1976) showed that a horizontally varying wind stress causes a mass transport 9 0 deg. to the right of the stress in the Ekman surface layer. This surface layer div- ergence is accompanied by an upward displacement of the iso- pycnals or upwelling of cccler water from below which tends to suppress the turbulent mixing process. During the relax- ation of the wind, these isopycnals are displaced downward , which contributes to the creaticn of large scale inertial- internal waves at the base of the mixed layer. Curing the passage of hurricane Belle (Mayer et al., 1981), ocean current and temperature measurements were acquired on the continental shelf of the Middle Atlantic Bight. Their analyses indicated that most of the inertial- internal wave energy was ccntained in the first mode at the 19 deeper sites (water depth of 70 m) and in a heavily damped second mode at the shallow sites (water depth of 50 m) . The variability in the modes was not attributed to the spatial variability of the wind stress. The bottom slope and mean velocity fields significantly altered the oceanic response to hurricane Eelle. A mixed layer model (Garwood, 1977) was embedded into a multi-layer, primitive equation, ocean circulation model (GCM) by Adamec et al., (1980). Using this GCM model, Hop- kins (1982) simulated the baroclinic response to a forcing pattern similar to hurricane Frederic and compared the results to the data collected in the wake of Frederic (Shay and Tamul, 1980). The model predicted the inertial response in the mixed layer quite well. However, the predicted ocean current response in the subsurface layers was less energetic than the observations, and decayed much too fast. These discrepancies between the model simulations and observations are due in part to the imposition of the rigid lid conditior at the surface, which eliminates the barotropic mode. Fur- thermore, these model simulations did not include topogra- phy, even though the observations are from a region where the bottom topography is rugged and the ocean depth is less than 1 km deep. 20 Recently, ocean measurements were made during the passage cf hurricane Alien in the western Gulf of Mexico (Brocks, 1983) . The spatial scales of the inertial wave were different from the 'site D1 observations, as shown in Table I. The vertical scales in the ocean are generally much greater under hurricane forcing than during the passage of a cold front, while the converse is true for the horizon- tal scales cf motion. The vertical group velocity, which depends on tcth the wave and the Brunt-Vaisala freguencies, for the Allen observations exceeded the group velocity from the 'site D' observations. This estimate from the Allen observations is misleading due tc the lack of upper thermo- cline and mixed layer data. Moreover, the wind speed is much greater during the passage cf a hurricane than a fron- tal passage. Therefore, the amount of turbulent mixing in the mixed layer and inertial wave excitation should be much more rapid during a hurricane. Other fea-ures measured in the Allen response included topographical dependence, and the vertical phase locking during the first few inertial pericd (IP) fcllcwing the storm. The baroclinic response of the ocean to a hurricane was modeled by (Price, 1983) using a multi-level, inviscid 21 model. The scales and energetics of the simulated inertial response were similar to mixed layer data collected by a NOAA data tuoy during the passage of hurricane Eloise. The horizontal and vertical scales, as given in Table I, were quite large in comparison to the thickness of the thermo- cline (200 m) . The maxiiium energy of the inertial-internal wave motion was predicted to occur at a distance of twice the radius of maximum winds, hereafter referred to as the maximum wind regime, which was roughly 80 km for Eloise. Further, the rate of vertical energy propagation was large in comparison to both the Allen and 'site D1 observations, and accounted for the depletion of energy from the mixed layer. Other mechanisms, such as KH instability (Pollard et al., 1973 ; Garwood, 1977) and turbulence (Bell, 1978), act to remove energy from the mixed layer. However, these phe- nomena were neglected as mixing was not explicitly included in the model formulation. Greatbatch (1983) modeled the nonlinear response of the ocean to a moving stcrm. The major results were that the transition between upwelling and downwelling zones in the oscillatory wake was rapid, and that nonlinearities account for the displacement of the maximum response to the right of the storm track. 22 III. DATA AND ANALYTICAL METHODS A. CURRENT METEB DATA Ten Aanderaa RCM-5 current meters were deployed on three moored taut-wire arrays ir depths ranging from 100 to 470 m in the northern Gulf of Mexico. These current meters sam- pled ccean current speed, direction and temperatures at 10 minute intervals. Two current meter arrays (CMA2, and 3) were deployed on adjacent sides of the DeSoto Canyon where bathymetric ccntcurs converge to form the head of the canyon (see Big. 1) . The other icoring was deployed closer to the coast in a depth of about 100 m of water where the isobaths are nearly parallel to the coast. A synopsis is given in Table II cf the storm period observations, which extends from a few days prior to hurricane passage (12 September) to the end of the deployment period (mid-October) . The guality cf the data is generally good; however, the lengths of the time series are not all the same. For instance, the Savonius rctors were eventually lost from all current meters in the mixed layer due to the large current speeds. These large mixed layer current speeds are not 23 corrected fcr rctor pumping, which is a function of the mooring design as well as the large direction vane of the Aanderaa current meters (Pofonoff and Ercan, 1967). in addition to the rotor problems, temperature records from all mixed layer current meters were unrecoverable because the ocean temperature exceeded the threshold temperature of the thermistors of 21.5 °C . Sea surface temperatures 325 km south-southeast of the mooring sites, as measured by NCAA Buoy 42003, reached 28.8<>C (Johnson and Renwick, 1981) . B. STORM TRACK The stcrm track is based on the best position data on hurricane Frederic's movement through the Gulf of Mexico (Hebert, 1979) . Because of the intensity of Frederic, recon- naissance aircraft constantly monitored the storm. Frederic increased to maximum strength, maximum winds or minimum barometric pressure, 80 to 130 km west of the CMA sites about 2100 GM1 12 September (see Fig. 1). A visible photo- graph of Frederic from a GOES satellite at 2001 GMT 12 September clearly delineates a well developed eye of about 40 tc 50 km in diameter (Figure 2) . The translational speed of the hurricane at this time was about 7 to 7.5 m/s as it approached the Gulf Coast, which was much larger than the 24 internal wave phase speed. Thus, an inertio-gravity wave response is expected in the wake of hurricane Frederic (Geisler, 1970). TABLE II A Synopsis of Stcrm Period Observations Meter Depth (1) Record Length Start Time (GMT) End Time (GMT) Variables CMA1 21 49 64 92* 0 22.21 16. "6 20.79 No hurricane data 0040 5 Sep. 0600 27 Sep. u,v 0040 5 Sep. 1920 21 Sep. u,v,T 0040 5 Sep. 2040 25 Sep. u,v,T CMA2 19 24.67 0040 5 Sep. 179 22.63 1920 7 Sep. 3 24 24.16 0040 5 Se p. 1650 29 Sep, 0100 30 Sep 1900 29 Sep, u,v u,v,T u,v,T CMA3 21 24.91 0100 2 Sep 251 36.00 0100 2 Se p 437 35.69 0100 2 Sep 457 51. S4 O1C0 2 Se p 2200 0100 1730 0000 2 6 Sep 8 Oct 7 Oct 4 Oct u,v u,v,T u,v,T u,vrT * time clock synchronization problems u,v = horizontal velocity components T = temperature GMT = Greenwich Mean Time 25 Figure 2. Satellite Photograph cf Frederic at 200 1 GMT 12 y Sept. 1979. Visual satellite imagery is courtesy of NOAA/NESDIS. 26 C. BIND 'FIELD Wind field data were obtained from shore stations at Mobile, Alabama and Pensacola, Florida. These records indi- cated that the surface wind speed did not exceed 40 m/s. However, ether reports suggested that wind speeds ranged between 48 to 58 m/s as Frederic made landfall (Hebert, 1979). These discrepancies are attributed zo local boundary effects and the non-representativeness of wind data col- lected in the coastal region. Marine winds were also meas- ured by NOAA Euoy 42003 (Jchnson and Renwick, 198 1). Wind speeds at this location never exceeded 35 m/s. The eye clearly passed over the bucy as indicated by the minimum in the wind field as the direction changed from 40 to 200 ° True (Fig. 3), with a corresponding decrease in pressure to 959 millibars (mb). D. VERTICAL TEMEERATURE GRADIENTS The AXBT data collected by the National Hurricane Research Division, as reported by Black (1983), are used for the computation of the Br UEt-Vai sala frequency. These data were collected after the passage of Frederic in the area of current meter array deployments near the DeSoto Canyon (see Fig. 1). 27 254 255 256 I I I I I I I I I I I I I I I I I I I I I M I 15 18 21 0 3 6 9 12 IS 256 254 255 JULIAN DAY/TIME (GMT) Figure 3. Wind Speed and Direction Measured by NOAA Buoy 42003. The upper panel represents the observed wind speed time series where the abscissa depicts time in Julian Days starting Dn 1200 GMT 11 Sept. to 1500 GMT 13 Sept. 1979. Ths lower panel represents the concurrent wind direction time series. The hatched area depicts the period when the data buoy was in the eye of the hurricane (Johnson and Renwick, 1981). 28 The vertical density profiles are computed at 10 m intervals using the AXBT data and climatological data from NAVOCEANO thrcugh the equation of state P = pQ(l-a(T - Tq)) 9 (1) where p is the density, T is the observed temperature, T0 and p0 are the reference temperature and density respec- tively frcn. climatology, and a is the thermal expansion coefficient, which is taken to be 0.0002/°C . Some error is expected due tc the neglect of the salinity term. However, an examination of the clicatological T-S diagram, for the DeSoto Canyon area indicates that density variations are due to temperature rather thai salinity effects. The Brunt-Vai- sala, N2 , frequency is computed using a centered finite difference frcm the expression *T2 g A p p Az / ^- > o where g is the acceleraticc due to gravity andAp/Az is the vertical density gradient. Since the AXBT data only extended to about 250 m, it was necessary -o extrapolate the vertical temperature gradient to 470 m. The temperature is 29 assumed to decrease uniformly at a rata of 0.0311 oc/m from 250 to 470 m (Fig. 4) • • The corresponding uniform increase in the density produces a constant Erunt-Vaisala profile. T (°C) 7.5 10.0 12.5 15.0 17.5 20.0 22.5 25.0 27.5 I ' 1 ' 1 ' 1 ' 1 ' 1 ' 1 ' 1 ' 1 N (cph) 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 Figure 4. Vertical Temperature and Brunt- Vaisala Frequency Profiles. The abscissa depicts the Brunt-vaisala frequency (solid) and temperature (dashed) from Black (1983) . E. TEMPORAL AND SPATIAL SCALES The lccal variations in the bottom topography at the DeSotc Canyon (see Fig. 1) are appproximately parallel and 30 normal to the coast at CM£2 and CMA3. For example, north- south flow at CMA2 is in the cress-shelf direction, but it is in the alcng-shelf direction at CMA3. Henceforth, a Cartesian coordinate system is used to represent the hori- zontal velocity components in the x and y directions, where motion to the east and the north corresponds to +u and + v current velocities, respectively. The fundamental time scale used in this thesis is an inertial pericc, which is equal to 24.10, 24.25 and 24.47 h at CMA 1,2 and 3 respectively. In the following calcula- tions, the Inertial Period (IP) will be set equal to 24.4 h. The intrinsic length scales are either the radius of maximum winds, approximately 40 ka; for Frederic, or the Rossby radii of deformation, which are given in Table III for the baro- tropic and first three barcclinic modes. These deformation radii are based en th€ inertio-g ravity wave phase speed com- puted from the Sturm-Liouville problem, which will be dis- cussed in a later section. For convenience, the scales in the x and y directions will be referenced as cross and along-track scales respectively, because the storm track was nearly perpendicular to the coastline. 31 TABLE III Rossby Radii of Deformation Phase Mode Speed No. (m/s) D ef ormati Radii (Jsi) 950. 56. 30. 18. on 0 68.0 1 4.0 2 2.2 3 1.3 F. COMPUTATION CF SPECTRA The energy and cross spectra are computed in the follow- ing manner: (1) the mean is computed and is extracted from -he time series; (2) a Tukey data window is applied to the data; (3) the data are transformed into frequency space via the Fast Fourier Transform (FFT) ; and (4) the transformed data are spectrally averaged over bandwidths which vary in width with frequency, and are closely tied to the computa- tion of confidence levels. The number of data points trans- formed by the FFT used in the spectrua analyses does not have to be proportional to 2n , where n is an integer. The general FFT is advantageous in the analysis of events because of their transient nature. However, in the analysis of the spectra, equal numbers of data points are used to avoid leakage and smearing of energy from adjacent frequency 32 bands (Otnes and Enochson, 1978) . Finally, the higher fre- quency energy abcve the Nyquist frequency, 1/(2 A *) where ^ t is the sampling interval, is eliminated to prevent ali- asing of the spectra. The HKZ is decomposed into the clockwise (CW) and coun- terclockwise (CCW) rotating components using the rotary spectrum methcd outlined in Gonella (1972) and Mooers (1972). The main results of this analysis are the rotary spectrum, ellipse stability and orientation, and rotary coefficient. The orientation of the ellipse indicates the direction of the horizontal phase velocity, while the sta- bility indicates the variability in the ellipse orientation. A high stability index indicates that the wave motion is anisctropic, or that the horizontal direction of phase pro- pagation is unidirectional. Conversely, a low index of sta- bility implies that the direction of phase propagation is randcm or isotropic. The rotary coefficient indicates the amount of confidence that can be placed in the rotary spec- tral estimate and the rotational component that dominates the rotary spectrum. For instance, a +1.00 implies that the moticn is polarized in the CM direction. A more rigorous mathematical treatment is given by Gonella (1972) and Mooers (1972). 33 G. COMPLEX DEMODULATION ' A useful method for determining inert ial oscillations in a time series of ocean current measurements is through the use cf the complex demodulation method (Perkins, 1970). The mathematical details are given in Appendix A. This method can te implemented either ty linear filtering or by perform- ing a least sguares harmcnic analysis on the time series. The linear filtering method is preferred because data points are lost whenever harmonic analyses are used to form the time-varying amplitudes and phases of frequency depen- dent motions. This is especially useful in the analysis of events when all data points are needed to resolve the tran- sient motions. A major disadvantage of the linear filtering method is that it only applies to discrete frequencies, whereas the least squares method can treat all frequencies within the constraints of the time serias. Initially, the data are band-pass filtered between 18 and 30 h to limit the frequency content in the time series to near-inertial motions. These filtered data are multi- plied by the trigonometric arguments of the inertial fre- quency. At this point, some higher frequency motions have been introduced into the time series (Appendix A) and these 34 are removed ty low-pass filtering the data at 22 h. The resultant time series is manipulated to form the instantane- ous amplitudes and phases cf the inertial motion. The final set of time-varying amplitudes and phases provide a local harmonic analysis, or more explicitly, a time series of spectral estimates for inertial motion. The filter used in the demodulation was a Lanzcos square taper window with 24 1 weights (Fig . 5) . The band-pass filtered, mixed layer velocity components at CMA3 are excellent examples for using complex demodula- tion (Fig, 6). As the current velocities increase in response to the storm, the amplitude of the forced inertial wave increases as well. The inertial wave amplitudes modu- late the increases and decreases in the mixed layer veloci- ties. After the maximum velocities occur, roughly 2 IP following hurricane passage, the strength of the currents decreases linearly with e-folding scales of about 4 IP. This type of e-folding behavior is very common in amplitude- modulated signals. 35 Figure 5. Frequency Filters. represented by the filter response is shewn Response for the Band-pass and Low-pass The tand-pass filter response is by the solid line while the low-pass by the dotted line. Time (Julian Days) Figure 6. Band-Eass Filtered Mixed layer Component at CMA3. 7 Velocity 36 IV. STORM PERIOD In the following analyses, the data are divided into two periods: 1) the storm period observations as previously described in Table II; and 2) the quiescent period which encompasses the period frcm the start of the data to a few days prior tc the storm. This chapter focuses on the iner- tial, subinertial and superinertial response in terms of energetics and scales of notion to hurricane Frederic. The quiescent period observations and results are given in Appendix B. For convenience, the terms shallow and deep will refer tc the regicns where CMA1 and CMA2,3 were deployed. A. INERTIflL RESPONSE Hurricane Frederic passed about 80 km west of the DeSoto Canyon at 2100 GMT 12 September (JD 255) 1979. The current speeds began to increase at 070 0 GMT or approximately 14 h prior to the time of closest approach. The currents in the mixed layer at CMA2 increased to a maximum within an IP fol- lowing the passage of Frederic (Fig. 7) . The storm caused the horizontal velocity to oscillate with periods of about 24.3 h and to rotate in a CW direction, which indicates the 37 presence of inertial waves in the mixed layer. These waves then decreased in amplitude over e-folding scales of 4 IP. The storm also caused the u-component to flow in the oppo- site direction from that cf the pre-storm flew, whereas the v-component increased markedly in amplitude. The near-bct- tom current speeds (shown later) began to increase within 3 h after the mixed layer response. c c 5 o <" o 120 IX 60 G0- 40- 20 0- -20- -40- -60- AAA^^' 2+e /WwV^ 256 260 264 Time (Julian Days) 26? 272 Figure 7. Mixed Layer Velocity Components at CMA2. The ordinate depicts the magnitude of the horizontal velocity components in cm/s for the east-west current (upper) and north-south current (lower) . The hatched area depicts the period of hurricane passage. 38 In the mixed layer at CMA3 , the storm amplified the pre-storm currents to a maximum of 135 cm/s (Fig. 8). As at CMA2, this increase in the currents is a manifestation of inertial wave excitation. However, the inertial response at CMA3 was different from CMA2 in that the kinetic energy of the waves was larger and persisted longer (approximately 2 IP) . The energy of these waves then decreased over e-fold- ing scales of 4 IP. The increase in near-bottom current (shown later) occurred within 4.5 h following the mixed layer response. These horizontal differences of the inertial response in the mixed layer between CMA2 and CMA3 are observed in the amplitude changes of the HKE (Fig. 9). The initial increase in kinetic energy at the two arrays was similar. The maxi- mum in kinetic energy occurred first at CMA2 and led the mixed layer maximum at CMA3 by about 3 h. The differences in HKE levels were quite pronounced. For example, the mixed layer HKE level was roughly 60 % of that found in the mixed layer at CKA3, even though CMA2 was closer to the eye of the storm. This large difference over a distance of 34 km may be due in part tc the stronger mean current at CMA2, which may have reduced the amplitude of the inertial waves. The 39 Time (Julian Days) Figure 8. Sinilar to Fig. 7, except at CMA3. direction of the mean flow and the horizontal phase propaga- tion cf the inertial waves (shown below) roughly coincided, which suggests that there may have been some interaction between them. Following these maxima, the energy decayed with an e-folding scale cf 4 IP at both arrays. Forced inertial waves in the mixed layer that vary in space and time are one mechanism to drive inertio-gravity waves in the subsurface layers (Krauss, 1972a, b). 40 m U n a) >-i td T3 0) X . ST3 0) a>x: J= Ui +j M 3 en •H 0s2 Q91X 009T 0921 0001 09i (1) (d U HO) C J3 0) T3H U Q) 0) X M-l •H t*M e— :H T3T3 <:-h aci-i a; UU£ OV -MS idMlf) d)-M W^U t^-M-H x a. u> u> "4-1 J- 13 0-*-1 GJ_t •H^ •ri,CX WiHO eu-t-> u> •H UJ (0*U EH-'^C-*-' f^ 0) M 3 •H OS-iJ 005J 0521 0001 05i 005 052 0 47 Although the inertial peak was prominent, the variability in spectral shapes was due tc the differences in the background continuum cf internal waves which varies with depth (Fu, 1980) and bandwidth averaging. The HKE spectra are decomposed into CW and CCW rotating components to clarify the rotational characterise ics of the inertial wave motion. The spectral peak near the inertial freguency was shifted about 3-6% higher than the inertial freguency and it dominated the CW spectra calculated from the mixed layer velocity measurements at CMA3 (Fig. 14) . The rotary spectrum estimates for the inertial period motions are given in Table IV. The CW, inertial HKE vari- ability estimates are generally 1 to 2 orders of magnitude larger than the CCW rotating motion. In comparison, the subinertial (periods of 3 IP) and superinertial (periods of 12 h) CW spectral estimates are more than an order of magni- tude less than the inertial motion (Tables VII and VIII) . At all levels at the deep region, the positive rotary coefficients approached unity, which indicates that the waves were polarized in the CW direction (Table IV) . The direction of tie semi-major axes of the inertial ellipses at CMA2 rotated CW with depth except near the bottom. The 48 o^ ptp CjO- >> 1 CflU dp-: CD T~ OS f \A o5 CD- CD a' O- 21M 251M 437M 457M K. M, CI I "I — 1IM -3 10 Figure 13 10 10 eye les/hr Normalized HKB Spectra from CMA3 at 21 m (solid) , 251 m (dashed) , U37 m ( dashed-dot ted) and U57 m (dotted). K, represents the diurnal tide frequency, f the mertial frequency and Mx the semi-diurral tide frequency. H9 is the phase difference between similar velocity components (Pollard, 1980). This analysis is restricted to the CMA2 and 3 due to the absence of an energetic inertial response at CMA1 (Table V) . Esti- mates of vertical scales are more accurate than the horizon- tal scale estimates because the vertical spacing of the current meters below the nixed layer is not the same at the two arrays. These scales calculated from the entire storm time series will be compared belcw with the scales calcu- lated from the instantaneous phases. The vertical cross-spectral analyses (Table V) indicate that the inertial response in the mixed layer led the near- bottom response. The normalized cross-spectrum variance between the vertical levels was about a factor of two greater at CMA3 than at CMS2. The vertical scales computed from (3) are about 360 and 540 m at CMA2,3 respectively. 52 That is, the inertial motion had scales of the order of the water depths at both arrays and was coherent at the 95 % confidence level. TABLE V Vertical and Horizontal Coherencies at the Inertial Frequency Nor nalized Cross-Spectra Squared Phase Arrajj Variables Variance/c ch Coherence (j|£3«) Vertical 2 u(s) ,u(b) 3. 2x10+* 2 v (s) ,v % (0 Q C o cd CO "^ cv 3 OS S lO jr* c\2 r-« CO lO C\J 000T 09^ 009 092 (sja^a^) apses iftSiiaq ■lO C\2 CO hifi c\? lO m C\2 x: a>+j -p c a) . sue o * 4-j cum jOcr a) f/Jts 3 U (0 ems o n 14-1 CN m»4-t U +■> co-p (didc -c a) UJ UL.C u> o (0 3 s uoo to CU i rO-P+J U G-H •*i (0 U 4J-M O M 0J»-1 W C ci> >-H > m (P M •H fa 55 along-track (north-south) than in the cross-track (east- west) directions. These estimates of scales may be somewhat misleading because of the instrument depth difference (135 m) , although they agree well with mixed layer estimates. The instantaneous phase differences between the mixed layer currents indicate that the horizontal length scales were set bj the storm initially (Fig. 16). The scales decreased rapidly and approached constant values of 60 and 100 km in the east-west and north-south directions, respec- tively. These scales in the mixed layer are roughly 1-2 times the baroclinic Rossby radius of deformation for the first mode (56 km). The estimates of the horizontal scales of inertic-gravity waves in the thermocline are computed from the fcottcn: and mid-depth instantaneous phase differ- ences in the 0-velocity ccmponent at CMA2 and CMA3 , respec- tively (Fig. 17) . The east-west and north-south scales in the thermocline immediately after hurricane passage are roughly the same as in the mixed layer. After 4 IP, the scales revert to the pre-storm scales of 25 and 50 km in the east-west and north-south directions, respectively. Seme caution is advised in accepting these estimates due to the large depth difference of 73 m. In a later section, it is 56 suggested that the energy was contained primarily in the barotropic and first two karoclinic modes. Hence, a depxh difference of 13 m may be acceptable within the deeper ther- mocline because of the relatively large displacement of the isotherms, the large vertical scales, and the amount of energy contained in the barotropic mode. The estimates of vertical energy propagation speed are computed only en the basis of the lag times between the sur- face and subsurface response (Table VI). In the first case, the response is felt at the bottom of the water column within 3 to 4.5 h after increases in the mixed layer are detected. Tie initial response corresponds to a vertical scale of roughly 900 m (Fig. 15). This lag of 3 to 4.5 h results in a vertical energy propagation speed of 2.6 km/d which is roughly 1 to 2 orders of magnitude greater than the values calculated by Brooks (1983) and Price (1983). The 10 IP lag in the second maximum observed at 251 m represents a vertical energy propagation speed of 26 m/d when the verti- cal scale is about 500 m. This vertical propagation of energy agrees guite well with the estimates derived from the Allen observations. However, there is also a second maximum observed near the bottom atout 6 IP following the storm. 57 021 •H +>cs 0J4J PS -p tn (0 CO) EU'CI'H in 04J^i cp * cuw o J3X! S CLOd) U >i d Ul Oi-O •H 3 >*d o-PPHid w o>-h a) (U+JU 4-> o-p o)sa w a > a> .H-P UKN +'13+' M £5.h on it) O C«aS NO) axi) a+j m WO) O 33 a) (ti ctyt o o -P-Q d)«— m rtJ Mwl T3 a u 0/ _c o> •H4M*s: M w W (O-H B O <0 v£> 0> u 3 •H (sj9}9iiion>l) 9I^°S m^u^l 58 Consequently, the vertical (CgE ) and horizontal (Cg^ ) group velocities are calculated from the following expres- sions (Brocks, 1983) Cgz " " m ~2 tan 9 7 (4) 0 n _ a N2 2a CgH " k ^2 tan 9 7 <5) where cr is the observed frequency, m and k are wavenumbers in the vertical and horizontal cross-track directions, and tan0 - k/m. The wavenumbers tn and k are equal to 2 n /L2and 2 tt /Irrespectively, where Lz and LH represent the vertical and horizontal scales as estimated previously. The assigned values to the variables in (4) and (5) are a ■ 7. 54 x 10-s rad/s, N2 = 3. 30 x 1C-s rad2/sz, m = 6.98 x 1 0"3 rad/m, m s 1.16 x 10-2 rad/m, k ■ 1.05 x 10-* rad/m. The initial vertical group velocity is 1.25 km/d as com- puted from the dispersion relation (4) with an initial scale of 900 m (Table VI) . This estimate is about half of 59 09T v m ■V 'US w 0)CU (U.C l3t0 C! t UU 3 > o c=r o M sajtN en a t-tum i u •i-t-x! OJ3 cue* a -H O »-H c ,-i— .a. 0 0)0)3(1) N^3 > T3 •H4J «- O G.CCN M 33 O-tJ^iO M 3 IT •H 921 001 9-L 09 98 (sj9^9uionx) 9I130S tft^usq 60 TABLE VI Comparison cf Vertical and Horizontal Group Velocity. Vertical Brooks * Price * Scales (1983) (1983) 900 m 500 m Vertical Group Velocity Lag Time 2.6 0.03 Disoersion 1.3 C.27 .06 .160 (km/d) Horizontal Group Velocity Dispersion 80. 30. 23. 86. (km/d) * From Brooks (1983) the cne computed based on the initial lag time and is still an order of magnitude larger than the estimates computed from the Eloise simulations. After a few IP, the vertical group velocity is about 0.27 km/d, which is of the order of the Eloise estimates cf abcut 0. 16 km/d (Price, 1983) . Fur- thermore, this estimate is about one order of magnitude larger than the value of 0.02 km/d that Brooks (1983) esti- mated from the Allen observations. Clearly, the initial storm forcing causes energy to propagate vertically at a significantly higher rate than after a few IP following the storm. 61 The horizontal propagation of energy is a much more rapid than the vertical propagation (Table VI). The group velocity estimate for the Frederic observations is initially 80 km/d; and it decreases to about 30 km/d a few IP follow- ing the storm. This difference is due to the tan2 depen- dence on m as previously defined. The initial horizontal group velocity estimate agrees with the Price (1983) esti- mate whereas the lower value agrees with the Brooks (1983) estimate. The horizontal phase speeds of the inertio -grav- ity waves in the wake of Frederic are about 70 cm/s with a vertical phase speed of rcughly 1 cm/s. In summary, inertial waves maintained a constant ampli- tude over 1 to 2 IP following the passage of Frederic. These oscillations in the mixed layer then relaxed over an e-folding scale of about 4 IP. In the subsurface layers, the forced inertio-gravity waves decayed over similar time scales after the appearance of a secondary maximum. The lag between the initial impulse of energy and the second subsur- face maximum varied over depth. For example, this period was about 10 IP at 251 m, but it was only 6 IP at 457 m. These lag values define a modulation envelope which evolves in time. If vertical propagation of energy from the mixed 62 layer alone accounts for these secondary maxima, why do the maxima occur later in the deep thermocline than near the bottom ? Cther mechanisms besides dispersion of energy from the mixed layer must be responsible. For example, nonlinear resonant interactions among the inertio-gravity waves could be responsible (McComas and Bretherton , 1977). Analyses of these interactions are beyced the scope of the present work. The initial vertical scales of the inertio-gravity waves in the wake of Frederic are approximately egual to the Eloise and Allen estimates. Hcwever, after a few IP the vertical scale decreased to about 500 ra or approximately the water depth. The horizontal scales estimated from the Fred- eric ebservatiens differ greatly from the Allen data and Eloise simulations. The horizontal scales during Allen were four times the mixed layer horizontal scales observed immed- iately after the passage of Frederic. This difference increased to a factcr of about eight approximately 4 IP after Frederic. These discrepancies may be due to the rug- ged bcttom topography in the DeS oto Canyon region. B. SUBINEBTIAI RESPONSE The forcing by hurricare Frederic also altered the "mean flow" and the lower frequency motions (periods greater than 63 3 IP) . Th€ variability and response of the mean flow to the forcing is depicted in progressive vector diagrams (PVD) . The ocean current data were low-pass filtered with a half power point at 26 h to isolate the forced response on the lower freguency variability. The mean flow in the nixed layer at CMA2 was directed towards the east prior to the storm (Fig. 18a). During the passage of hurricane Frederic near JD 256, there was only a small deflection in the mean flow. The subsequent horizon- tal displacements indicate that the mean flow was about 26 km/d, or almost twice the mean current during the pre-stcrm period. The mean flow at 179 m (not shown) at CMA2 also increased towards the east. The bcttca PVD at CMA2 indicates that the mean flow at that level was mere or less constant (Fig. 18b). However, the direction of this flow was opposite to that observed in the mixed layer. Consequently, the bottom PVD was nearly a mirrcr image of the surface PVD, although the bottom flow was slower than that of tte mixed layer. There was also a superposed lower frequency motion of about 3 I? near the bottom, particularly during the period between the initial and secondary maxima of HKE (first 6 IP in Fig. 9). The 64 lower frequency oscillations may be due to coastally trapped waves, which were enhanced by the storm, since waves of the same period were observed during the quiescent period (Appendix E) . These waves can be altered by both bottom topography and continuous density stratification (Wang and Mooers, 1976). A study of the mechanisms by which the mean flow is altered by these waves and the rugged terrain is also beyond the scope of this thesis. At CMA3, the surface mean flow was southward pricr to the storm (Fig. 19a). At the time of storm passage, near JD 256, the surface mean flew was deflected outward from the storm center. Between JD 257 to 269, the mean flow acceler- ated to about 13 km/d and then decelerated to about 6 km/d towards the south. Prior to the storm, the mean flow near the tcttom at CMA3 was ncrthward following the local bottom topography (Fig. 19b) . There was an eastward (up the canyon slope) deflection late on JD 256. After completing a loop, the mean flow veered towards the northwest (down the canyon slope). Between JD 254 to 263, the mean flow near the bot- tom was considerably faster than either earlier or later times. Again, the mean flow near the bottom was reversed relative to that in the mixed layer. 65 Scale: \- ^64 Figure 18 70 km Lew-pass Filtered PV C at CMA2. The * represents a time interval of 48 h starting on JD 248 for a) mixed layer (19 m) and b) bottom Layer (324 m) . Thermocline temperatures in the wake of Frederic varied spatially as well temporally. There were sizeable subiner- tial oscillations as well as near-inertial oscillations in the temperature signal (Fig. 20) . These temperature varia- tions were typically 3.0°C at both array sites and were superposed on mean temperatures that range between 13.0 to 66 J Scale: \ Figure 19. Lew-pass Filtered PV E at CMA3. The * represents a time interval of 48 h starting on JD 245 for a) mixed layer (21 m) and b) bottom Layer (457 m) . 14.5°C in the thermocline. At CMA3r the temperature at 251 m started to increase about 0600 GMT on JD 255 and continued to increase over the next 15 h due to the strong northward current. The temperatures then decreased by 2<>C due to a strong southward current. Inert ial fluctuations of abcut 3oc persisted ever the subsequent 8 IP . After this time, 67 the variations underwent a slight increase in frequency which corresponded to about the time of the secondary maxi- mum in the horizontal velocity components. For the remain- der of the record, the temperature variations were superposed on a stationary mean. Eottom temperature variability at CMA2 is a clear exam- ple of advective processes that are associated with lower frequency motions (Fig. 21). On JD 254 (11 September), a temperature decrease of 1.3°C over a 16 h period marked the beginning of the forcing period as cool water was advected from the DeSoto Canyon. The temperature then increased by 3.70 c over the next 25 h as warm water was advected by the southward velocity component which was downslope in this part of the canyon. As the winds relaxed, cooler water was again advected into the area. Around JD 260, the tempera- ture dropped fcy 4°C ever an 11 hour period. This drastic change in temperature was due to the advection of cold water by these subinertial waves of 3 IP period. From the orien- tation of the ellipses (Table VII) , the direction of the wave propagation was northward (upslope, away from the DeSotc Canyon) . These waves set up cross-shelf oscillations (relative to CMA2) in the north-south direction. After 68 0'9T m z: u e o M •H fcH(N o CM W M 3 en •H Cm OH OCT OET 69 these large variations, near-inertial oscillations in the temperature were superposed on a gradual warming trend. TABLE VII Normalized Botary Spectrum Analysis at Subinertial Frequency Rotary Coef f . Depth (m) Normalized Rotary Spectra CW CCW Elli Stab. ps e Dir. (Dea.) 324 457 4.5x10*3 3.0x10+3 4.7x10+3 1.0x10 + 2 0.65 0.50 10 10 ♦ 0.20 ♦ 0.98 The potential energy spectra ware computed using the temperature time series and the vertical temperature gradi- ents from the AXET data collected by Black (1983) (Appendix A) . The mcst prcminent peak in the potential energy spectra at CMA3 is near the inertial frequency (Fig. 22) . The most energetic peaks at the semi-diurnal and subinertial frequen- cies are near the bottom. Potential energy variability in the bottom layers at CMA2 (not shown) is significant at the 95 % confidence levels only for periods of about 3 IP. From rotary spectrum analyses at periods of 3 IP near the bottom: (1) the CW and CCW rotating components at CMA2 were nearly equal, with just a slight tendency towards CW polarization; and (2) the motion at CMA3 was polarized in a 70 O'tl cm ro (d S3 u e o M-i W Q) •H M 0) CO 0) M s B EH U •H Cm 0"ZI (5) an oot 06 ajn^jaduiaj, 71 10 cycles/hr Figure 22 Potential Energy Spectra from CMA3 isotherm displacements observed at (solid) , 437 b (dashed) represents the diurnal inertia! frequency and frequency. and tide for 251 457 m (dot frequency, the m :ed) . f the K< M- the semi-diurnal tide 72 CW sense (Table VII) . Furthermore, the phase propagation for these lcnger period waves was northward, and anisotropic as indicated ry the ellipse orientation and stability. The cross-spectrum between the temperature and north- south velocity records at the bottom was also dominated by the peak at about 3 IP (Fig. 23). The corresponding phase and coherence estimates are 100 deg. and 0.98 respectively, with the velocity leading the temperature signal. Hence, these lower frequency oscillations in the cross-shelf direc- tion at CMA2 were properly oriented for advecting the temp- erature pattern. These cross-shelf oscillations were responsible for the variability in the bottom temperature 4 IP after the passage of Frederic. C. SOPERINERTIAL RESPONSE As Frederic passed within 130 lem to the west of the CMA1 , the dominant response was a significant increase in the westward current (Fig. 24) . This increase was due to convergence of flow on the right side of the storm. 3y con- trast, there was a divergence of flow on the left side of the hurricane, where water was transported off-shelf by the winds. For example, a negative tide was recorded at Biloxi, Mississippi which was indicative of the left side of the 73 storm. The north-south current only increased by about 20 cm/s, whereas the east-west current increased to 80 cm/s. After the hurricane passage, near-inertial waves persisted over the subsequent 3 IP. The post-storm semi-diurnal tidal currents were mere energetic than during the quiescent period and were superposed on a non-stationary trend. At the intermediate depth at CMA1, the current speed increased to 80 cm/s during the period of strong forcing, as the semi-diurnal tidal currents were enhanced. The presence of internal tides at the semi-diurnal fre- quency is indicated by the HK E spectra during the storm period at CMA1 (Table VIII). The spectral estimates in the semi-diurnal frequency band are almost equal to the inertial period estimates. Baines (1973) shows that one necessary condition for the presence of internal tides is a marked increase in the semi-diurnal tidal currents. The observed increase in the semi-diurnal tidal currents should be equal to or greater than the inertial effects. The storm-spectral estimates at the semi-diurnal frequency exceed those from the guiescent period by mere than order of magnitude (Appen- dix B) . a barotropic tide propagating over the shelf break generates internal tides cf semi-diurnal tidal period 74 o_ cycles/hr Figure 23. Cross-Spectrum Between V-Velocity and Temperature from CMA2 at 324 m. Kv represents the diurnal tide frequency, f the inertia! frequency and Mt the semi-diurnal tide frequency. 75 40-| *J • c a; 20 1 G*-v 2.5; o- c M -20- d H ■ i ° -40- D . -<30- -eo- 60 -i *j C 0] 40- o 9 <~%. 20- o ,- o - o > -20-1 -40 24£ 256 264 Time (Julian Days) Figure 24. Mid-depth Velocity Components from CMA1. The ordinate depicts magnitude in cm/s of the east- west currents (upper) and north-south currents (lower) . (Prinsenberg and Rattray, 1975 ; Barbee et al., 1975 ; Tcr- grimson and Hickey, 1979). Hence, an increased barotropic tide associated with the storm surge propagating over rugged bottom topography should enhance the internal tides as mani- fested in the semi-diurnal tidal currents. 76 TABLE VIII Normalized HKE Spectra at Inertial and Semi-diurnal Frequency Normalized Meter No. of Data Frequency HKE Spectra Depth Points Resolution f M2 (§) (£2h) (SI^s) 2/cph 3200 0 .0018 2.0x10*3 1.0x10+3 2410 0.0025 3.0x10*3 6.0x10+2 3000 0 .0020 1.4x10+3 3. 2x10+3 CMA1 49 64 92* * instrumentation problems with time clock M2 is the semi-diurnal tide frequency band f is the ineitial/diurnal tide frequency band The dominance of internal tides on the shelf at CMA1 can be explained ty considering the bottom slope in the region. If the bottom slope is greater than the internal wave char- acteristic, then the rays will not be admitted to the shelf, but instead will be reflected into the oceans interior (Bar- bee et al, 1S75 ; LeBlond and Mysak,1978 ; Torgrimson and Hickey, 1979). The critical slope for the internal wave characteristic is given by: ** = (°2 - f2)1/2 (6) dy \2 _ c2} 1 K } where f is the local Coriclis parameter, which is equal to 7.18 x 10~s rad/s near the DeSotc Canyon. The remainder of 77 the variables have been defined above. The slope of the in ertio-gravity wave characteristic is about 4 x 10-3, whereas in the direction of the canyon axis (45 deg.) the bottom slope is 6 x 10~3 . Therefore, it is clear that the inertio-gravity waves will be reflected seaward as the slope of the characteristic is less than the bottom slope in the DeSctc Canyon region. Inertial oscillations will be locally generated on the shelf but they will only persist for a few IP following the storm. The internal tides, however, will be admitted to the shelf region because the slope of the semi-diurnal tide characteristic is 2 x 10~2, which is almost an order of magnitude larger than the bottom slope. The semi-diurnal tides were also enhanced at CMA2 and 3, but they were much less energetic than the inertial waves (Table IX) . The CW polarized, semi-diurnal tidal variabil- ity was more than an crder of magnitude greater than the CCW variability estimates, except in the bottom layers at CMA3. The directions of the semi-major axes of these waves changed from 120 tc 20 deg. with depth at CMA2, whereas the opposite occurred at CKA3, where the direction changed from 30 to 50 deg. Hence, the waves were anisotropic and exhibited a strong north-south component of flow at CMA2 and to the northeast at CMA3. 78 TABLE IX Normalized Bctary Spectra at the Semi-diurnal Frequency. Normalized Rotary Spectra Ellipse Eepth CW CCW Dir. Rotary (m) (cm/s) 2/cph Stab. (Peg.) Coef f . a) CMA2 19 1.3x10+3 U.3X10+2 0.80 179 5.0x10+2 9.0X10+1 0.73 324 4.3x10+2 1.2x10 + 2 0.87 120 + 0.30 60 + 0.72 20 + 0.41 k) £MA3 21 1.0x10+* 5.5x1 0+2 0.58 30 + 0.55 251 6.7x10+2 1.0x1 0+2 0.92 90 + 0.70 437 3.2x10+2 1. 3x10+2 0.40 40 + 0.40 457 2.0x10+2 1.8x1 0+2 0.15 50 + 0.04 In summary, the superinert ial response in the wake of Frederic was dominated by the enhanced semi-diurnal tidal current at all three arrays. The semi-diurnal tidal cur- rents were markedly more energetic than the inertial cur- rents on the continental shelf at CMA1 , and agrees with internal tide theory (Baines, 1973) . The storm surge, which is associated with the passage cf the storm, increases the barotropic tide. As this increased barotropic tide propa- gates over rugged bottom terrain in a stratified ocean, the increase in the semi-diurnal tidal currents are manifesta- tions of internal tides (Earbee et al, 1975 ; Prinsenberg and Rattray, 1S75 ; Torgrimson and Hickey , 1979). Internal 79 tides can propagate freely up the continental slope and onto the shelf, whereas inertio-gravity waves generated off the slope are reflected towards the interior of the ocean. The rotational characteristics of the semi-diurnal -ides, as well as the direction of phase propagation of these waves, are consistent with these theories. 80 V. NORMAL MODJS A. THEORY The Sturm-Licuville picblem is solved using the vertical profile of Brunt-Vaisala frequencies and linear wave theory. The vertical eigenvalues cf the horizontal velocity are then used to ottain a least sguares fit to the Fourier coeffi- cient time series computed from the demodulation of the cur- rent meter records. The amount of variability associated with each mode is determined to assist in understanding the modal structure of the inertio- gravity wave response gener- ated fcy the passage of a hurricane. The problem is formulated by following the work of Fjeldstad (1958) and making the following assumptions: f plane; continuous stratification; hydrostatic balance with basic state at rest; incompressible; inviscid; and flat tottcm. 81 The governing vertical structure equation for linear, iner- tio-gravity waves is given by 2 £-| + y2$(z)W = 0 -> (7) dz where W(z) gives the vertical structure of the inertio-grav- ity waver /f=l5''oa-fx , $(z)=N2- 0-2 , and K2 = fc2 + 12, and k and 1 are the horizontal wave number vectors in the x and y directions, respectively. Furthermore, equation (7) is a second order, homogeneous, ordinary differential equation. The specification of the problem is completed by imposing the following boundary conditions: f£ - guW = 0 z=0 7 (8) dz ' W = 0 z=-d . (9) The surface boundary condition (8) is a dynamic boundary condition which states that at the sea surface pressure is continuous across the interface. Boundary condition (9) specifies no normal flow through the bottom. The solution to equations (7-9), which define a well known Sturm-Liou- ville problem, yields the structure of the vertical velocity for constant N, and, is given by Wn(z) . sin(2J£) n (10) 82 where the mode number n = 0,1,2.... The horizontal velocity eigenfunct ions 0 (z) are given by Un(z) ~ cos(H^) 7 (11) which is the vertical derivative of wn (z) . The Sturm-Licuville problem is solved numerically by the predictor-corrector fourth-order method modified by Hamming (Gerald, 1980). This methcd computes a new vector from four preceding values. A fourth-order Runge-Kutta method is used for the adjustment of the initial vertical increments and the computations of starting values. The resulting amplitudes of the horizontal velocity eigenfunctions fcr modes 0 through 3 are shown in Fig. 25 . The number of zero crossings equals the mode number. Most of the vertical structure lies in the upper thermocline regicn near the base cf the mixed layer. B. DATA ANALYSIS The horizontal eigenfunctions are then fitted in a least squares sense to the time-varying amplitudes of the horizon- tal velocities. The mathematical details are given in Appendix A with the final matrix equations (16-18). The time evolution cf the north-south velocity coefficients at 83 o d -6.0 -4.0 -2.0 Uh(z) 0.0 2.0 4.0 6.0 o o o o o o 02 PL, o CO o d ID CO O d o q d to o d o lO LEGEND MODE 0 MODE 1 "MODE'S" "MODE "3 Figure 25. Horizontal Velocity Eigenf unctions at 29.33 N and §7.11 w. The abscissa depicts the amplitude of the horizontal velocity eigenf unction for mode 0 (solid) , mode 1 (dashed) , mode 2 (dotted) ana mode 3 (dashed-dctted) . 8U CM A3 indicates that the storm excited modes 0,1, and 3 with a damped mode 2 (Fig. 26) . The amplitudes of modes 0 and 1 reached a peak on JD 256 and then decreased. By JD 262, the barotropic mode horizontal speed decreased to nearly a con- stant value of about 10 cm/s. After JD 262, modes 0 and 1 oscillated over the remainder of the record with a period of about 5 IP. This behavior of the modss in Fig. 26 indi- cates that a modulation occurred between inertial wave modes 0 and 1. More importantly, these modes define a modulation envelope which is consistent with the amplitude variations of the currents below the mixed layer. At CMA2, the time-varying coefficients of the modes are markedly different in their behavior (Fig. 27). The spatial variability in the modal structure between CMA2 and CMA3 is associated with the differences in the mean flow, which is topographically controlled. Initially, all of the modes were excited, with a fairly energetic lode 3. The baro- tropic coefficient decreased to about 2 cm/s while the mode 1 amplitude approached zero near JD 252. After JD 262, modes 0 and. 1 were in phase and defined the limits of the coefficients. 35 ►J-25 OOiO OS CO 10 CM CD ^~-^ CM 00 >> CO Q e CO - — h ,-^~ r- 3 10 1-3 CM *— ' 0 s • *H E- CO 10 CM 05 CM OC 02 01 0 01- (s/uio) apniTjduiy r 02- oc- CM T3 Pi (0 a^ o ow r\i en «« a> Err-! u o a (0 «. WO • COO 0) W 0) •H (t)4-> 0«O-P •H^O <4-l «0 4-l«- I d) IT3 O <1)iS (0 •H »w COf*") 0)«-tQ) > CO W O >wa M •H 86 Q!QiQ OOiO O CD CN C\J cm cO Q -M (0 a CO a. CO 3 o "-8 X s>— ^ 0) 0) CO cm i 02 OT 0 01- (s/uio) 9pn^Tjdiny \-\o cm CN •H Cm 18 0) s to CN CP M 3 a> •H Cm 02- 0C- o CM 87 The mccal current time series are re-computed using the coefficients from the least sguares fit and the eigenfunc- tions for each cf the current meter depths at CMA3. For example, the reconstructed time series are compared in Fig. 28 with the actual demodulated current time series at 251 m . Recall that a secondary maximum occurred at this site about 10 IF following the passage of Frederic (Fig. 10). The reconstructed amplitude cf the barotropic mode is greater than the actual v component of the data between JD 255 and 262 (Fig. 28). However, summing modes 0,1 and 2 accounts for most cf the observed variability, with the barotropic mode the most energetic. Over the remainder of the record, JD 262 to 270, most of the observed variability can be described cnly by modes 0 and 2 . The contribution of the barotropic mode was significant and consistent with earlier estimates that the vertical scales were of the order of the water depth. The strength of the barotropic mode continued for the entire time series of the modal coefficients at CM A3. Estimates of the modal contributions to the observed near-inertial variability are given in Table X for both the storm and poststorm periods at CMA2,3. These estimates are based on the expression 88 X 2 £(V. - V. ) s = 1. - J - , (12) j J where V; and VjM are the observed and modal scalar components of the horizontal velocity. The index m depicts the summa- tion of the modes, for example, 0, 0+1, and 0+1+2, and inde j repesents a summation ever the number of observations in the period. The storm period is redefined to start at the time of hurricane passage and continue for 7 IP. The post- storm period starts where the storm period ends and contin- ues for roughly 8 IP until the end of record. The mixed layer variability at CMA2 is dominated by the barotropic mode during the storm period. The contribution of the barctrcpic mode tc the observed horizontal current variability decreases with depth, but by including the baroclinic modes most of the near-inertial variance can be described by modes 0,1 and 2. The only exception is at 179 m. The first three modes contribute only 54 and 37% to the observed horizontal current variance in the east-west and north-south directions, respectively. Furthermore, the u-velccity component of the barotropic mode exceeds the 39 TABLE X Variance cf the Normal Modes at CMA2 and CMA3. Depth Velocity (m) Component Period CMA2 19 u s 19 u ps 19 v s 19 v ps 179 u s 179 u ps 179 v s 179 v ps 321 u s 324 u ps 324 v s 324 v ps CM A3 Modes 0 0 + 1 0+1 + 2 (%) (%) (%) 68 78 100 87 91 100 69 78 100 92 93 100 53 50 54 98 98 95 35 31 37 97 98 95 * 50 67 97 98 99 45 72 85 97 99 99 69 97 100 73 86 100 68 98 100 * 50 100 55 81 98 95 96 99 58 84 98 94 95 99 73 68 97 95 94 91 64 78 94 94 94 92 22 82 85 79 74 99 21 89 84 81 66 99 21 u s 21 u ps 21 v s 21 v ps 251 u s 251 u ps 251 v s 251 v ps 437 u s 437 u ps 437 v s 437 v ps 457 u s 457. u ps 457 v s 457 v ps s: storm period ( number of observations = 1008) ps: Dost-stcrm period (number of observations = 1152) *: an overestimation of the variance (explained below) observations by a considerable amount and causes equation (12) to be less than zero (see Fig. 27). This behavior could be attributed to the bottom boundary layer where infi- nite energies occur when the slope of the internal wave characteristic approaches the bottom slope (Prinsenberg and Rattray, 1975) . Most of the observed variance during the 90 post-storm period is attributed to the barotropic mode. The modes are much more well-behaved during this period, which is presumably a manifestation of free waves. The modes of the horizontal velocity are well-behaved during the stcrm period at CMA3, except near the bottom where only 20 to 22% of the observed variance can be attrib- uted to the barctropic mode. This is probably due to bottom boundary effects. At the other depths, the barotropic mode explains 55 tc 73% of the variance, with most of the vari- ance accounted for by the summation of the barotropic and first two barcclinic modes. During the post-storm period, the observations of the nixed layer velocity are less than that of the barctropic mcde in the north-south direction, which causes the value to be less than zero. The reason for this large discrepancy is due to the e-folding scale of the horizontal velocity in the mixed layer, the persistence of the coefficient cf the barotropic mode (see Fig. 26) , and the radiation of inertial waves away from the storm track. Otherwise, 73 tc 95% cf tbe observed horizontal current var- iance can be attributed to the barotropic mode, and adding the first two barcclinic mcdes accounts for more than 90% of the variance. 91 Hopkins (1982) modeled only the baroclinic response to hurricane passage and he found that the inertio-gravity waves decayed toe rapidly in the thermocline and bottom lay- ers to account for the vertical structure of the observed variability. Thus, a model with only the baroclinic modes could not explain the secondary maximum observed in these layers. Some caution has to applied in the interpretation of the normal modes for several reasons. First, the formulation of the Sturm-Liouville problem assumes that the vertical struc- ture consists of a standing wave for the N2 = constant case. From basic physics, a standing wave can be decomposed into two waves of equal amplitudes propagating in opposite direc- tions, which assumas no phase propagation. Over the first 7 to 1 0 IP, there is a downward propagation of energy, as well as upward propagation of phase, which violates the standing wave argument. Secondly, a flat bottom ocean is assumed for the application of normal mode theory. rha validity of this assumption crucially depends on the ratio between the slope of the internal wave characteristic to the bottom slope. For example, as this ratic approaches unity, the energy den- sities in the bottom boundary layer approaches infinity. At 92 245 250 255 260 265 270 20 O GO > £ > 10- o 20 O ^-v 0 JQ 10 0 20 >> -J-i • iw4 a ^ o ^ 10 > s 1 ° > 245 i i i MODE 0 MODE 0+1 250 255 260 Time (Julian Days) 265 270 Figure 28. V Component Time Series at CMA3 for the observed time series (solid) and the reconstructed time series (dotted) for a) mode 0 , b) mode 0+1 , c) mode 0+1+2 . 93 that point, dissipation becomes important and must ba included in the model (Prinsenberg and Rattray, 1975). Con- sequently, the assumption of in viscid dynamics is no longer valid. Despite all these approximations, the normal mode analysis fits the observations quite well. Even the reso- nance or nodulation envelope which occurred at CMA3 about 10 IP following the storm passage, is represented quite well by modes 0, 1 and 2. 9U VI . CO NCL 0 SI CNS The ocean response near the DeSoto Canyon to hurricane Frederic was dominated by the excitation of inertial waves in the mixed layer and inertio- gravity waves in the thermo- cline. These waves are not admitted onto the shelf because the slope cf the characteristics of internal wave motion is less than the bottom slope (Barbee et. al, 1975 ; LeBlond and Mysak, 1S78 ; Torgrimson and Hickey, 1979). The major features cf the forced inertio-gravity waves observed at CMA2 and 3 are: • initially, the vertical propagation of energy was of the order cf 1. km/d as the forcing was felt throughout the water column within 3 to 4.5 H following the'pas- sage cf Frederic; • the initial horizontal propagation of energy was about 80 km/d with a phase speed of 70 cm/s ; • the waves were anisotropic at all depths and had an upward propagation of phase; • maximum HKE in the mixed layer occurred at the fringe of the naximum wind regime with a second maximum in the thermocline 6 to 10 IE after the storm; • the EKE decayed on an e-folding scale of about U IP after the maximum in the mixed layer and the second maximum in the thermccline; • the vertical propagation after about 6 IP is of the order of 102 m/d with a corresponding horizontal propa- gation of energy of 30 km/d; • the mixed layer horizontal scales were 60 and 100 km in the cress and along-track directions, respectively, while the scales in the thermocline were 25 and 50 km in the cress and along-track directions; 95 • the vertical scales were cf the order of the water depth which suggests that the barotropic mode was important; • most cf the observed variabilty was in the barotropic mode, with some contributions from modes 1 and 2; together they defined a modulation envelope in the deep thermccline at CMA3; • the modal coefficients varied in space as well as time due tc the rugged bcttcm topography and mean currents. The mean flew was topcgrapically controlled and reversed direction frcm surfaca tc bottom. The forced response in the mixed layer caused the mean flow to increase at both CMA2 and CMA3. The mean flow below the mixed layer was also changed by the storm. Furthermore, this mean flow was influenced ty subinertial variability (periods of 3 IP) and was linked to the advecticn of cooler water cross-shelf rel- ative to CEA2. The semi-diurnal tidal currents increased in magnitude and varied spatially in response to the passage of hurricane Frederic. The near dominance of the semi-diurnal tidal cur- rents over the inertial motions at CMA1 was consistent with internal tide theories. The semi-diurnal tides were admit- ted to the shelf region because the slope of the internal wave characteristic exceeded the critical bottom slope 96 (Baines , 1973 ; Prinsenberg and Rattrayr 1975 ; Torgrimson and Hickey, 1979) . In a brcader context, the oceanic response to hurricane Frederic was a geostrophic adjustment problem, which had both transient and steady state components. After the tran- sient, inertic-gra vity waves propagated away from the storm track, a gecstrophically balanced ridge and current system remained under the storm track. That is, The steady state currents were due to the balance between the horizontal pressure gradients and the Coriclis force (Geisler, 1970). Although some adjustments to the mean flow are described, additional study is reguired to completely understand the interelat icnships between the forced mean flow, the forced wave field and the bottom topography. 97 APPENDIX A A. COMPLEX DEMODULATION The motivation for complex demodulation is to isolate the carrier wave of a certain frequency, which must be known a priori , and to form amplitudes and phases of the modulated signal. A linear filtering approach is used to form the instantaneous amplitudes and phases in contrast to the har- monic analysis method. Essentially, the amplitude and phase time series represents a lccal harmonic analysis rather than discrete estimates averaged over continuous periods (Otnes and Enochscn, 1978) . A band-limited time series is multiplied by the trigono- metric arguments of the carrier wave. The resultant time series is then lew-pass filtered to eliminate some high fre- quency noise that is generated. The cosine and sine coeffi- cients are ther combined tc form the amplitude and phase of the modulated wave. Consider the time series: x(i) = A is the phase , A is the amplitude of the wave, and i 98 represents a time index. Both the phase and amplitude vary as functions cf time. Multiply equation (1) by the sine and cosine arguments of the carrier frequency fc , xs (i)=A (i)iccs ((2irfciAt) + 0 (i) )1 sin (2 tt fc iAt ) xc (i)=A 6 E* (1U) 3 ^OjE^Z;) =C,]TE,(Zj) E^(z;) ♦CaVE2(z,) S3(Z;) i«l •=! 3 ♦ C3yE3(Zi ) E3(z,) (15) By inspection, let a*M=/JWZi > E* o d- 5 V (TJ m *2 X3 /*™v (J w 2? >^ s S «5 o Q u C w 03 0) •»■* •H __ . »-t U co H V 33 CO ?3 fa 9^ ?3 111 50 k tvv Figure 32. Kid-depth PVE at CMA1. The * represents a time span of 48 h starting on JD 207 at 49 m. due to internal clock problems. For example, large increases in the velocity occurred on JD 252 as opposed to 256. Hence, the data were not included. The HKE spectra for the 21, 49, and 64 m depths are shown in Table XUI for bcth the inertial and semi-diurnal frequency tands. The inertial/ diurnal tidal kinetic energy 112 in the surface layer was two orders of magnitude larger than in the subsurface layers. The semi-diurnal tidal motion decreased from the surface to 49 m and then increased with depth such that the HKE near the bottom almost equaled that near the surface. TABLE XIII Normalized HKE Spectral Estimates at 95% Confidence Normalized Meter No. of Data Frequency HKE Spectra Depth Points Resolution f M2 (J) 2£b (cm/s)+2/cph CMAJ 21 5825 0.001 1.4x10+* 1.5x10+2 49 5825 0.001 1.4x10+3 4.3x10+1 64 5825 0.001 5.0x10+2 8.0x10+1 92* 5825 0.001 1.5x10+2 1.0x10+2 * instrumentation problems with time clock M2 is the seui-diurnal tide frequency band f is the inertial/diurnal tide frequency band Temperature data at bcth 21 and 49 m are not available because of the thermistor problem. The temperatures ranged from 18.0QC to the thermistor cutoff of 21.5 °C at 49 m. Temperature variations were about 0.5OC over an IP, although most of the variability was associated with some non-sta- tionary trends and lower frequency motions similar to those observed in the currents. Jus- prior to the storm, tempera- 113 tures decreased to a relative minimum as cold water was advected fcy tfce mean flow. In summary, based on the observations during the quies- cent period, the mean flew strongly depends on the bottom topography. Inertial band motion is evident at all depths during the quiescent period. Seme of this motion is due to freely propagating inertial waves while the remainder is part of the forced diurnal tides. Internal tides, as well as inertial oscillations, are part of the freely propagating internal wave continuum below the surface layer. The effect of the topography on the mean flow and the wave motions fur- ther complicates the dynamics of the region and the resul- tant circulation . Longer period waves (2.5 to 3 day pericd) set up north-south oscillations that advect cooler water from the DeSoto Canycn onto the shelf region. Super- inertial frequency motions are influenced by the diurnal and semi-diurnal tides and contribute to the variability. 1 14 LIST CF REFERENCES Adamec, D . - R. L. Elsberry, R. W. Garwood Jr. , and R. L. Haney, 1980, An embedded mxed-layer ocean circulation bo del, Dy_n. oj Atmos. and Oceans , 6, 6 9-9 6. Baines, P. G., 1S71,The reflection of internal/inertial waves from bumpy surfaces, J. Fluid Mech. , 46, 273-291. Baines, P. G., 1973, The generation of internal tides by flat-tump topography, Deec-Sea Res., 20, 179-205. Barbee, W. D., J. G. Dvorski, J. D. Irish, L. H. Larson, and M. Rattray Jr., 1975, Measurement of internal waves of tidal frequency near a coastal boundary, J. Geo£h_y_s, Res., 15, 1965-1974. Bell, T. H., 1978. Radiation damping of inertial oscillations in the upper ocean, J. Fluid Mech. , 88, 28 9-205. ~ Bendat, J. S., and A. G. Eiersol , 1971, Random Data: analysis measurement procedures, Wiley-Interscience, New York, 407 pp. Black, P. G., 1963, Ocean temperature changes induced by tropical cyclones, Ph. D. disseration, The Penr.sylavania State Univ., 278pp. Brooks, D., 1983, The wake of hurricane Allen in the western Gulf of Mexico, J. Phl§. Cceanogr. . J3, 117-129. Elsberry, R. L., T. Fraim, and R. Trapnell, 1976, A mixed layer model of the ocean thermal response to hurricanes J. Geo£hvs. Res., 8J, 1153-1162. Fjeldstad, J. E. , 1958, Ocean current as an initial problem, <£eo.pJvy,sj.s.ke Pub., 20(7), 1-24. Fofonoff, N. P. and Y. Ercan, 1967, Response characteristics of a savonius iotor current meter. Woods Hole Oceanographic Institution, WHCI ref. no. 67-33, Fu, L-L., 1981, Observations an models of inertial waves in the deep ocean, Rev. Geophys. Space Phv^s. , 19(1), 141-170. Garwood, R. W., Jr., 1977, An oceanic mixed-layer model capable of simulating cyclic states, J. Phys. Oceancgr., 7, 45 5-468. * J " — *— *- Geisler, J. E., 1970, Linear theory on the response of a two layer ocean to moving hurricane, Geophys. Fluid Dy_n. , 115 Gerald, C. F., 1980, Applied Numerical Analysis, Addison- Wesley, Readirg, Massachusetts, 518 pp. Gonella. J.r 1972, A rotary component method for analyzing meteorological and oceanographic time series, Deep-Sea Res., 19, 633-846. Greatbatch, R. J., 1983, Cn the response of xhe ocean to a moving storm: the nonlinear dynamics, J. Phys. Oceanoqr., 13, 357-3 67. " — *- *~ Hebert, P., 1979, Preliminary report: hurricane Frederic Aug 29 - Sept. 14,1979, NOAA National Hurricane Center, Miami, Florida. Hopkins, C, 1982, Ocean response to hurricane forcing, M. S. Thesis, Naval Postgraduate School, Monterey, California, 89 pp. Johnson, A. and S. Renwick, 1981, Buoy observations during the passage of hurricane Frederic 1979, Data Report, NOAA Data Buoy Office, NSTL Station, Mississippi. Kase, R. H. and D. J. Olbers, 1979, Wind driven inertial waves observed during phase III of GATE, Deep-sea Res., 261, 19 1-216. c Krauss, H.f 1972, Mind generated internal waves and inertial pericd motion, Dtsch. H^drc^r. Z., 25 , 241-250. Krauss, W.p 1S76a, On currents, internal and inertial waves in a stratified ccean due to variable winds, Part 1, Dtsch. Sldrc.gr. Z., 29, 87-96. Krauss, W.f 1976a, On currents, internal and iner-ial waves in a stratified ocean due to variable winds, Dtsch. Hydroqr. Z. , ^9, 1* 1- 1 35. LeBlcnd, P. H. and L. A. Mysak, 1978, Waves in the ocean, Elsevier, Amsterdam, 638 pp. Leipper, D., 1967, Observed ocean conditions and hurricane Hilda, 1964, J. Atmos. Sci., 24, 182-196. Mayer, D.f M. C. Mofjeld, and K. D. Leaman, 1981, Near- inertial internal waves or the outer shelf in the middle Atlantic Bight in the wake of hurricane Belle, J. Phys. Oceanocjr. , jj[, 8 6-106. McComas, C. H. and F. P. Eretherton, 1977, Resonant interactions of oceanic internal waves, J. Gegphys. Res., 82, 1397-1412. "" 116 Mooersf C. N. K.f 1973r A technique fcr the cross spectrum analysis of complex valued time series with emphasis on properties of polarized components and rotational invariants, Dg€grSea Res. , 20, 1129-1141. 0* Brier., J. anc R. 0- Reid, 1967, The nonlinear response of a two layer taroclinic ocean to a stationary, axially- symetric hurricane, I, Upwellinq induced by momentum transfer, J. Atmcs. Sci. , 24, 208-215. Otnes, R. K., and L. Enochson, 1978, Applied Time Series Analysis, Wiley-Int erscience. New York, pp. 449. Perkins, H., 1970, Inertial oscillations in the Mediterranean Sea, Ph.D. dissertation, Massachusetts Institute of Technology, Woods hole Oceanographic Institution, 123 pp. Pollard, R. I., 1970, On the generation by winds of inertial waves in the ccean. Deep- Sea Res., 17, 795-812. Pollard, R. I., 1980, Properties of near-surface inertial oscillations, 0. Phvs. Oceanogr. , JO, 385-398. Pollard, R. T. and R. C. Millard, 1970, Comparison between observed and simulated wind generated inertial oscillations, Dee£ Sea Res., V7, 8 13-82 1. Pollard, R. T.- E. B. Rhines, and R. Thompson. 1973, The deepening cf tie wind mixed layer, Geophys. Fluid Dyn. , 3, 38i-a04. — *-*— L~ Price, J., 1983, Internal wave wake of a moving storm, Part I, scales, energy budget and observations, J. Phys. Oceanogr . , 12, 94 9-965. Prinsenberg, S. J., and M. Rattray, Jr.f 1975, Effects of continental slope and variable Brunt-Vaisala frequency on the coastal generation of internal tides, Deep-Sea Res. ,22, 25 1-263. Shay, L. K. and J. J. Tamul, 1980, Abstract: Current observations in the wake cf hurricane Frederic, EOS, vol. 61 (17), p. 256. Torgrimson, G. M., and B. Hickey , 1979, Barotropic and baroclinic tides over the continental slope and shelf off Oregon, J. Phv,s. Oceanogr . ,9. 94 5-96 1. Wang, Dong-Pina« and C. N. K. Mocers, 1976, Coastal trapped waves in a continuously stratified ocean, J. Phys. Oceanoar., 6, 6 5 3-863. "" 1 17 INITIAL DISTRIBUTION LIST No. Copies 1. Defense Technical Infcrmaticn Center Camercn Station Alexandria, VA 2 23 1 U 2. Library, Code 0142 Naval Postgraduate School Monterey, CA 93943 3. Professor Robert J. Renard, Cede 63Rd Department of Meteorology Naval Postgraduate School Monterey, CA 93943 4. Professor Christopher N. K. Mooers, Code 68Mr Department cf Oceanography Naval Postgraduate School Monterey, CA 93943 5. Professor Russell L. Elsberry, Code 63Es Department cf Meteorology Naval Postgraduate School Monterey, CA 93943 6. Dr. Andrew Willmott, Code 6 8W Department cf Mathematics University cf Exeter North Parfc Boad Exeter EX4 4QE England 7. Dr. William Hart, Code 8000 Naval Oceancgraphic Cffice NSTL Station Eay St. Louis, MS 39522 8. Mr. Lynn K. Shay, Code 63 Dept. of Meteorology Naval Postgraduate School Monterey, CA 93943 9. Director Naval Oceancgraphy Division Naval Observatory 34th and Massachusetts Avenue NW Washington, D.C. 203SO 10. Commander Naval Oceancgraphy Ccimand Central NSTL Station Eay ST. Louis, MS 39522 1 18 11. Commanding Officer Naval Oceanoaraphic Office NSTL Station Eay St. Louis, MS 39522 12. Commanding Officer Fleet Numerical Oceanography Center Monterey, CA 93940 13. Commanding Officer Naval Ocean Research and Development Activity NSTL Station Eay ST. Louis, MS 39522 14. Commanding Cfficer Naval Environmental Prediction Research Facilit )zl Monterey, CA 93940 15. Chairnan, Oceanography Department D.S. Naval Academy Annapolis, MD 21402 16. Chief of Naval Research 800 N. Quincy Street Arlington, VA 22217 17. Office of Naval Research (Cede 480) Naval Ocean Research andDev elepment Activity NSTL Station Eay ST. Louis, MS 39522 18. Scientific Liaiscn Office Office of Naval Research Scripps Institution of OceanograDhy La Jolla, CA 92037 19. Library Scripcps Institution cf Oceanography P.O. Eox 2367 La Jolla, CA 92037 20. Library Department cf Oceanography University of Washington Seattle, BA 98105 21. Library CICESE P.O. Eox 4803 San Ysidro, CA 92073 1 19 22. Library School of Oregon State Corvallis, OR 97331 Oceanoaraphy University 23. Library Coastal Studies Louisiana State Eaton Rouge, LA Institute University 70603 24 . Commander Cceanographic Systems Pacific Eox 1390 Fear Harrcr, HI S686C 25. Commander (Air-370) Naval Air Systems Command Washington, D. C. 20360 26. Chief, Ocean Services Division National Oceanic and Atmospheric Administration 8060 Thirteenth Street Silver Springs, MD 20910 120 Thesis S437365 c.l 207664 Shay- Observations of in- ertio-gravity waves in the wake of hurricane Frederic. 207664 Thesis S437365 Shay c.l Observations of in- ertio-gravity waves in the wake of hurricane Frederic.