<<

AUGUST 2003 KURAPOV ET AL. 1733

The M2 Internal off Oregon: Inferences from Data Assimilation

ALEXANDER L. KURAPOV,GARY D. EGBERT,J.S.ALLEN,ROBERT N. MILLER, SVETLANA Y. E ROFEEVA, AND P. M . K OSRO College of Oceanic and Atmospheric Sciences, Oregon State University, Corvallis, Oregon

(Manuscript received 8 May 2002, in ®nal form 21 January 2003)

ABSTRACT A linearized baroclinic, spectral-in-time tidal inverse model has been developed for assimilation of surface currents from coast-based high-frequency (HF) radars. Representer functions obtained as a part of the generalized inverse solution show that for superinertial ¯ows information from the surface velocity measurements propagates to depth along wave characteristics, allowing internal tidal ¯ows to be mapped throughout the . Application of the inverse model to a 38 km ϫ 57 km domain off the mid-Oregon coast, where data from two HF radar systems are available, provides a uniquely detailed picture of spatial and temporal variability of the

M2 internal tide in a coastal environment. Most baroclinic signal contained in the data comes from outside the computational domain, and so data assimilation (DA) is used to restore baroclinic currents at the open boundary (OB). Experiments with synthetic data demonstrate that the choice of the error covariance for the OB condition affects model performance. A covariance consistent with assumed dynamics is obtained by nesting, using representers computed in a larger domain. Harmonic analysis of currents from HF radars and an acoustic Doppler pro®ler (ADP) off Oregon for May±July 1998 reveals substantial intermittence of the internal tide, both in amplitude and phase. Assimilation of the surface current measurements captures the temporal variability and improves the ADP/solution rms difference. Despite signi®cant temporal variability, persistent features are found for the studied period; for instance, the dominant direction of baroclinic wave phase and energy propagation is always from the northwest. At the surface, baroclinic surface tidal currents (deviations from the depth-averaged current) can be 10 cm sϪ1, 2 times as large as the depth-averaged current. Barotropic-to-baroclinic energy conversion is generally weak within the model domain over the shelf but reaches 5 mW m Ϫ2 at times over the slopes of Stonewall Bank.

1. Introduction bottom intensi®cation of the baroclinic currents occurs. The most plausible site for generating energetic internal Internal are generated over sloping topography, tides in the coastal is the break, where the vertical component of the barotropic tidal where over the distance of several kilometers bathym- current forces oscillations in the density ®eld (Wunsch etry changes from supercritical (on the continental 1975; Baines 1982). The M 2 tide off Oregon is super- inertial such that energy propagates away slope) to subcritical (on the shelf) (Figs. 1, 2). from generation sites along wave characteristic surfaces, The Oregon coastal shelf is relatively narrow (30±50 or beams. The slope of the wave characteristics can be km wide) such that the internal tide propagating onshore estimated as remains energetic up to the coast, enhancing spatial and temporal variability in currents, mixing, and biological 2 2 1/2 2 2 Ϫ1/2 tan␸ ϭ (␻ Ϫ f )(N Ϫ ␻ ) , (1) productivity. The total M 2 tidal current in this region Ϫ1 where ␸ is the angle that the characteristics make with may reach 15 cm s or more, while maximum baro- the horizontal, ␻ is the tidal frequency, f is the inertial tropic (depth-averaged) currents are typically about 5 frequency, and N is the buoyancy frequency. Internal cm sϪ1 or less (Hayes and Halpern 1976; Torgrimson waves incident upon supercritical (where and Hickey 1979; Erofeeva et al. 2003). In coastal wa- the bottom slope is steeper than the wave characteristics) ters, internal tides show spatial variability on the scale re¯ect toward deeper water, while those incident upon of several kilometers and temporal variability on the subcritical bathymetry will propagate into shallower wa- scale of days. The ®rst-mode internal M2 horizontal ter. Where the bathymetric slope is close to tan␸, strong length scale, estimated as 2NH/␻, is 14 km in water of depth H ϭ 100 m, given a typical value of N ϭ 0.01 sϪ1. Topographic variation may introduce additional Corresponding author address: Dr. Alexander L. Kurapov, College of Oceanic and Atmospheric Sciences, Oregon State University, 104 length scales. One reason for high temporal variability Ocean Administration Bldg., Corvallis, OR 97331-5503. may be the spring±neap cycle (Petruncio et al. 1998). E-mail: [email protected] Other causes of variation may be changes in the strat-

᭧ 2003 American Meteorological Society 1734 JOURNAL OF PHYSICAL VOLUME 33

FIG. 1. Maps: (a) central part of the Oregon shelf. Contours are bathymetry every 50 m, but depths larger than 750 m are not contoured; shaded (white) areas are the regions where the slope is subcritical (supercritical) with respect to

the M2 internal tidal wave characteristics, corresponding to the horizontally uniform strati®cation shown in Fig. 2; stars denote locations of the two HF radars. (b) Computational domain for the representer analysis and inversion [its boundary is shown in (a) as the bold gray line]; contours are bathymetry, every 20 m; dots are the locations where HF surface current data are obtained; circles are the reduced basis representer locations (see section 3c); the star is the ADP mooring location. i®cation or in the low-frequency background currents, acteristics (Torgrimson and Hickey 1979). Intensive which may result from surface heat ¯ux or wind forcing wind-induced vertical mixing near the coast may weak- on the shelf. During the lifting of isopycnal en the baroclinic tidal current (Hayes and Halpern surfaces near the coast affects the angle of wave char- 1976). As a result, there are signi®cant seasonal vari- ations, with much weaker internal tides in winter (Er- ofeeva et al. 2003). Modeling internal tide variability on the shelf is dif- ®cult because of lack of information about surface heat ¯ux, wind forcing, and strati®cation on the scales of interest. The requirement for ®ne spatial resolution re- stricts the size of the computational domain, and the need for open boundary ¯ows contributes further to un- certainty. Observations of shelf currents may provide additional information about internal tides that can be used to constrain the model by means of data assimi- lation (DA). Since 1997, measurements of the surface currents on the mid-Oregon shelf have been available from two land-based high-frequency (HF) radars, over an area about 50 km ϫ 50 km, with resolution 1±2 km in space, and 20±60 min in time (Kosro et al. 1997). FIG. 2. Potential density␳ and corresponding buoyancy frequency In this paper, the focus is on semidiurnal tidal variations N used in computations: solid line is the mean of summer observa- during May±July 1998, when the HF surface data were tions, 1961±71, at a station 46 km offshore of Yaquina Head (Smith et al. 2000); the dashed line is a variant used to test solution sensitivity complemented by the currents from an acoustic Doppler to strati®cation (see section 7a). pro®ler (ADP) mooring, providing information about AUGUST 2003 KURAPOV ET AL. 1735 the ¯ow variability with depth. Oke et al. (2002) used 1998 is validated against the ADP, followed by analysis this dataset to show the value of assimilating HF radar of the internal tide on the Oregon shelf. Section 8 con- measurements into a model of wind-forced coastal cir- tains a summary. culation. For that study the data were low-pass ®ltered to eliminate tidal and inertial oscillations, and the focus 2. Data analysis was on subinertial time scales. Data assimilation is a way to estimate the state of the For the period of May±July 1998, time series of sur- ocean that best ®ts both the dynamics and the data (Ben- face currents on the mid-Oregon shelf are available from nett 1992). Since the data provide extra information for two HF radars (SeaSonde instrument systems, manu- the model, they may be utilized to correct the model factured by CODAR Ocean Sensors), installed at Ya- inputs in some way: to correct open boundary values, quina Head and Waldport (Fig. 1), as well as currents estimate model parameters, and/or recover errors in the from an ADP mooring installed 25 km off Yaquina dynamical equations. Optimal DA schemes require es- Head. The ADP, anchored at a depth of 80 m, provides timates of the model solution error covariance (Kurapov time series of velocity measurements in the water col- et al. 2002). Ideally a DA system for the coastal ocean umn at depths 12±68 m every 4 m. The ADP data were would model both tidal and wind-driven circulation and harmonically analyzed in a number of overlapping 2- be based on implementation of the fully nonlinear dy- week time windows such that variations associated with namical equations. However, before such a complicated the spring±neap tidal cycle are largely ®ltered out. Com- system can be built, it is appropriate to focus on some plex harmonic constants for eastward (u) and northward fundamental questions. Do measurements of surface (␷) components of velocity are computed in windows currents contain information about internal tides at centered on days 131±191, spaced 4 days apart. In the depth? If so, how can this surface information be pro- following these time-varying estimates are referred to jected downward in an optimal way? What are the prin- by the center day in the window. Substantial intermit- cipal features of the model solution error covariance in tence of the internal tide is seen in the tidal ellipses of the superinertial band? These questions are addressed the ADP currents (Fig. 3a). The ®rst baroclinic mode here with a model based on simpli®ed dynamics for remains dominant, but with amplitudes and phases vary- which a rigorous, variational, generalized inverse DA ing with time. During days 131±155 the estimated ADP method (GIM) is readily implemented (Bennett 1992, M2 baroclinic velocities are relatively low, with maxi- 2002). The DA model presented below is based on dy- mum currents about 5 cm sϪ1 both near the bottom and namics that describe linear harmonic oscillations of the the surface. Starting around day 159 the internal tide at strati®ed ocean with respect to a state of rest, on realistic the ADP location becomes much larger, reaching 9 cm bottom topography. Both generation and propagation of sϪ1 near the surface and 7 cm sϪ1 near the bottom. internal waves are essentially linear mechanisms, so ap- During days 159±187, the tide becomes surface inten- plication of this simple model to realistic data is prom- si®ed, as near-bottom currents gradually decrease to 4 ising. Our model does not include advection by mean cm sϪ1. The barotropic (depth averaged) current, with currents and accompanying spatial variability in the den- amplitude about 4 cm sϪ1, shows signi®cantly lower sity ®eld, which are limitations: wind-induced low-fre- variability than the deviations from the depth average quency baroclinic currents along the coast may be as (Erofeeva et al. 2003). Some variability in these esti- high as 0.5 m sϪ1, and their effect on the internal tide mates of barotropic tidal currents may be due to data cannot be addressed with our model. noise, the fact that the depth-averaged current is not The main goals of this study are to 1) learn about the exactly the barotropic current (separation of the ¯ow information content and utility of HF radar surface data into barotropic and baroclinic components on the slop- for modeling the superinertial internal tide, 2) demon- ping bathymetry is not obvious), and remnants of the strate the importance of a dynamically consistent inverse spring±neap cycle not entirely ®ltered as a result of data problem formulation at the open boundary, and 3) de- processing in short time windows. scribe the spatial and temporal variability of the internal A land-based HF radar infers the surface current from tide on the Oregon shelf at the dominant M 2 tidal fre- measurements of the radar backscatter from ocean quency. waves, considered with other information about the Section 2 provides an analysis of the ADP and HF waves (Kosro et al. 1997). Measured data were pro- radar data. The inverse model is described in section 3. cessed to radial vectors every 10 min, which were then In sections 4 and 5, the sensitivity of the inverse solution averaged over 1 h, providing maps of the radial com- to the open-boundary error covariance is studied in a ponent of the current with resolution 2 km radially and series of experiments with synthetic data. Section 6 dis- 5Њ azimuthally. The M2 harmonic tidal constants for the cusses the effect of assumed dynamics on representer radial components of surface velocity from the HF ra- functions that show the zones of in¯uence of the data dars were computed in a manner similar to that used and give the error covariance of the model solution. for the ADP data. To reduce error, harmonic constants Then, in section 7, realistic surface currents are assim- are estimated by ®tting measurements at four adjacent ilated and the series of inverse solutions for May±July locations of the data array for each HF radar system. 1736 JOURNAL OF VOLUME 33

FIG. 3. Horizontal M2 tidal current ellipses at the ADP data locations, decomposed into the depth average (shown above the zero depth) and deviations over the water column: (a) harmonically analyzed ADP data, (b) prior solution, and (c) inverse solution obtained by assimilating HF radar radials west of longitude 235.8Њ. Here and in Figs. 10 and 13 the ®lled ellipses correspond to CW vector rotation; the line from the ellipse center shows the velocity direction at zero phase. In the plot, northward velocity is directed up, and eastward velocity to the right.

Only harmonic constants with estimated mean least square (MLS) error Ͻ 5cmsϪ1 are assimilated. In the domain shown in Fig. 1b, the mean and median of MLS errors vary insigni®cantly with time, remaining close to 2cmsϪ1. During the study period the winds are mostly up- welling favorable (Fig. 4). This is expected to cause vertical mixing in a surface boundary layer and lifting of isopycnal surfaces closer to the coast. However, we lack speci®c empirical information about the density structure for this period. In the computations of the in- ternal tide the ambient potential density␳ (z) is assumed FIG. 4. Low-pass ®ltered wind stress, from wind measurements at Newport, Oregon (note that wind stress is shown here only for ref- to be horizontally uniform and temporally constant (Fig. erence, no wind forcing is applied to the model). 2). This approximation may be a serious source of error AUGUST 2003 KURAPOV ET AL. 1737

in our model, especially later in the study period since pation is a minor contributor to the momentum balance. during upwelling the pycnocline structure is gradually Since we do not solve the problem by time stepping, built up, and during short relaxation events (no wind) horizontal dissipation is not needed for computational the pycnocline does not return to horizontal. However, stability. Vertical diffusion of ␳ is not included in (4). we adopt this approximation since the inverse model is This simpli®es the energy balance and is consistent with run in a small domain, where the misspeci®ed open the assumption that vertical diffusion does not have ef- boundary baroclinic ¯ow is hypothesized to be a larger fect on the background␳ (z). source of error. Surface and bottom boundary conditions, at ␴ ϭ 0 and ␴ ϭϪ1, are uץ Inverse model .3 ϭ 0at␴ ϭ 0, (5) ␴ץ a. Model equations The model describes linear harmonic oscillations of w ϭ i␻␩ at ␴ ϭ 0, (6) the strati®ed ocean with respect to the state of rest. w ϭ 0at␴ ϭϪ1, (7) Boussinesq and hydrostatic approximations are made. uץ K State variables vary with time as exp(i␻t). The model M ϭ ru at ␴ ϭϪ1, (8) ␴ץ is written in horizontal Cartesian coordinates with x to H the east, y to the north. Linearized surface boundary where r(x, y) is the bottom drag coef®cient. conditions are applied at the mean surface level (z A depth-integrated continuity equation is obtained ϭ 0). The vertical coordinate is stretched into ␴ ϭ z/ with use of (3), (6), and (7): H(x, y), and so the model equations are (cf. Blumberg and Mellor 1987; Mellor 1998): 0 (Hu d␴ ϭ 0. (9 ´ ١ i␻␩ ϩ ͵ i␻Hu ϩ f ϫ Hu ϩ gH١␩ Ϫ1 ␳ At the side boundaries ⌫ the normal current u ´ n isץ ١H gH 2 0 :١␳ Ϫ ␴Ј d␴Ј speci®ed pointwise ϩ ␴Јץ ␳o ͵ H ␴ ΂΃ 0 at the coast (u u ´ n| ϭ (10ץ KM ץ Ϫϭ0, (2) ⌫ u at the open boundary. ␴ Ά OBץ ␴΂΃Hץ w Since horizontal advective terms are dropped from theץ Hu) ϭ 0, (3) equations, the problem with boundary condition (10) is)´ ١ ϩ ␴ formally well posed, in the sense that it has a uniqueץ solution which is stable to changes in inputs (Oliger and ץ (␳Hu) ϭ 0. (4)´ ١ i␻H␳ ϩ (␳w) ϩ ␴ SundstroÈm 1978; see also Bennett 1992, chapter 9). Theץ open boundary condition (10) is in general re¯ective. In (2)±(4), all the dynamic variables represent complex However, we are going to treat the boundary condition amplitudes: u(x, y, ␴) ϭ {u, ␷} is the horizontal velocity as imperfect and correct open boundary baroclinic ¯ux- vector, ␩(x, y) is the surface elevation, ␳(x, y, ␴)isthe es. This can be done in such a way to improve wave potential density perturbation about a horizontally uni- radiation out of the domain. form mean state␳␳ (z) ϭ (x, y, ␴), and w(x, y, ␴) is the Once the dynamics are de®ned, the GIM requires velocity component normal to the ␴-surfaces (w ϭ wcart speci®cation of a cost functional that is a weighted sum ١H, where wcart is the Cartesian vertical velocity of terms, each penalizing errors in the inputs such as ´ Ϫ ␴u component in the unstretched coordinates); the param- the model equations, boundary conditions, or data. This eters to be speci®ed are gravity g, the parameter can be minimized in an ef®cient way if the problem is

f ϭ const, the vertical viscosity KM(x, y, ␴), ba- separated into a series of so called adjoint and forward thymetry H(x, y) and the reference density ␳o ϭ const; problems (see Bennett 1992, 2002; Egbert et al. 1994). are the ``horizontal'' gradient and divergence Technically, there are two approaches to formulating the ´ ١ and ١ operators on a ␴ surface. generalized inverse problem that eventually differ in Vertical dissipation in (2) is the only sink of energy how the adjoint model is built (Sirkes and Tziperman in the model and it is a very simple and admittedly 1997). In the ®rst approach, the cost functional penalizes crude parameterization of turbulence associated with the errors in the continuous equations and boundary background ¯ows. A horizontal diffusion term is not conditions (2)±(10), and the ®nite-difference imple- included in (2), partly for a technical reason, to allow mentation is formulated independently for the forward the solution of the momentum equations in terms of and adjoint problems derived in continuous form. In the velocity locally for each vertical column (see appendix second approach, to be taken here, the cost functional A). In fact, for coastal applications, horizontal dissi- penalizes errors in the discretized dynamical equations 1738 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 33

and boundary conditions. The advantage of this ap- where the vector lk is the discrete data functional, the proach in our case is that the discrete adjoint solver will prime denotes the complex conjugate matrix transpose, be obtained as a matrix transpose of the forward solver, dk is the kth scalar observation value, ⑀dk is the data simplifying derivation, as well as computer coding, for error, and k ϭ 1,...,K. All the data can be combined representer calculation. in one matrix equation:

d ϭ LЈ␯ ϩ ⑀d. (18) b. Model discretization The inverse solution minimizes the cost function: The equations and boundary conditions are discreti- J(␯) ϭ (S␯ Ϫ f )Ј covϪ1(S␯ Ϫ f ) zed on a staggered Arakawa C grid, with details as in 00 Blumberg and Mellor (1987), and Mellor (1998). Then, Ϫ1 ϩ (d Ϫ LЈ␯)Ј covd (d Ϫ LЈ␯), (19) state variables at the points of the discrete grid are com- bined into vectors of unknowns u, w, ␩, and ␳, cor- where cov and covd are the Hermitian non-negative responding to continuous Hu, w, ␩, and ␳. Discretized de®nite covariance matrices of errors in the model equa- model equations can be written in matrix notation as tions and the data, respectively. These have to be spec- i®ed prior to inversion. In the statistical interpretation ⍀u ϩ G␩ ϩ P␳ ϭ fu, (11) of the GIM, the errors are treated as random vectors, so cov and cov , where angle brackets i␻␩ ϩ D u ϭ f , (12) ϭ͗⑀⑀Ј͘ d ϭ͗⑀d⑀Јd͘ a a denote ensemble average. It is also assumed that dy-

Ww ϩ Du ϭ fw, (13) namics and measurements are unbiased: ͗⑀͘ϭ0, and ͗⑀d͘ϭ0. B␳ ϩ Z␳␳w ϩ D u ϭ f␳, (14) In our case, the ®rst term in (19) incorporates pen- where (11) represents horizontal momentum equations alties on both the dynamical equations and boundary (2) combined with boundary conditions (5), (8), and conditions. If the errors in the dynamical equations and (10); (12) represents the depth-integrated continuity boundary conditions are assumed to be uncorrelated, equation (9); (13) is for the continuity equation (3) com- this term can be written as the sum of penalties on the bined with the bottom boundary condition (7); and (14) errors in these two sources. Note that, in general, dy- is the discrete representation of the equation for the namical and open boundary condition errors may be potential density (4). In (11)±(14), ⍀, G, and so on, are correlated (Bogden 2001). The inverse solution can be the coef®cient matrices resulting from discretization, expressed as and fu, fa, etc., are forcing vectors. If the state variables K and the forcing vectors are combined into ␯ ϭ ␯ ϩ b r , (20) inv 0 ͸ kk kϭ1   ufu where ␯ ϭ SϪ1f is the prior solution satisfying error- w f  0 0 ␯ ϭ and f ϭ w , (15) free equations (11)±(14), ␩ f    a Ϫ1 Ϫ1 rkkϭ Scov(SЈ) l (21) ␳ f␳ are representer vectors, coef®cients bk in (20) are ele- then Eqs. (11)±(14) can be formally written together as ments of vector b, of size K:

S␯ ϭ f, (16) Ϫ1 b ϭ (R ϩ covd)(d Ϫ LЈ␯ 0), (22) where S is the model operator (matrix). In general, the and R is the representer matrix with elements Rkj ϭ lЈkrj. forcing vector f can be written as f ϭ f 0 ϩ ⑀, where f 0 The representer shows the zone of in¯uence of each is the vector of boundary values and ⑀ represents errors observation in the model domain. It can be partitioned in the discretized equations and boundary conditions. into ®elds each providing correction to prior u, ␷,w,␩, Although in our study most of the elements of ⑀ are Ϫ1 and ␳. In (21), (SЈ) lk is the adjoint representer solution, always zero, we retain the general weak constraint for- which gives the sensitivity of the kth observation to mulation for future convenience (see section 3c). An changes in forcing and boundary conditions. It is mul- ef®cient approach that we utilize to solve (16) for an tiplied by the covariance, which usually has a smoothing arbitrary f is based on the direct factorization of the effect (see Egbert et al. 1994). Last, the ``forward'' prob- model operator (Egbert and Erofeeva 2002), as de- lem (21) is solved. Our solution method for (16) allows scribed in appendix A. rapid computation of a large number of representers (see appendix A). In our formulation, all the equations and c. Cost function and inverse solution boundary conditions are written as weak constraints. Strong constraint cases are recovered simply by setting Data to be assimilated can be written in the general the appropriate parts of cov to 0. form In the statistical interpretation of the GIM, repre-

dkkϭ lЈ␯ ϩ ⑀ dk, (17) senters give the prior error covariance for the model AUGUST 2003 KURAPOV ET AL. 1739

Ϫ4 2 Ϫ1 Ϫ2 2 Ϫ1 solution. Speci®cally, if e 0 ϭ ␯ 0 Ϫ ␯true is the error in 10 (m s ). Thus KM approaches 10 m s in the the prior solution, then upper 10 m (Wijesekera et al. 2003) and 10Ϫ4 m2 sϪ1 in deeper water [a minimum value based on the parame- r ϭ͗eeЈ͘l . (23) k 00 k terization of Pacanowski and Philander (1981)]. Increas- 1/2 Ϫ2 2 Ϫ1 In this interpretation, (Rjj) is the expected prior error ing KM to a constant value of 10 m s everywhere of the sampled quantity. Since the representer does not in the computational domain makes representer functions depend on the actual observed value, but only on the smoother and has only minor effects on the inverse so- form of lk and assumed cov [see (21)], the expected lution. The linear bottom friction coef®cient is a constant variability of the model solution can in principle be r ϭ 2.5 ϫ 10Ϫ4 msϪ1. Unless otherwise speci®ed, the assessed for any variable at any point. This property mean of summer observations, 1961±71, at a station 46 will be used to compute the open boundary covariance km offshore of Yaquina Head (Smith et al. 2000) de®nes by nesting (section 4). ␳(z) (the solid line in Fig. 2). In the area shown in Fig. 1b harmonic constants for The prior model is forced at the open boundary by the radial components of surface velocities from the HF barotropic (depth averaged) currents, taken from a radars are available at over 900 locations. Computation coastal-scale tidal inverse model in which TOPEX/Po- of the full set of representers would be costly. Conse- seidon altimetry data are assimilated into the shallow- quently, a reduced basis approach is taken (Egbert and water equations (Egbert and Erofeeva 2002). The same Erofeeva 2002), as detailed in appendix B. A set of 68 prior solution is used in each time window. The prior representers, chosen as a reduced basis for model cor- solution provides an accurate estimate of the barotropic rection, is computed for data functionals ˆlk correspond- semidiurnal ¯ow in the area (Erofeeva et al. 2003). ing to u and ␷ at a ®xed set of 34 surface locations (see However, deviations from the depth average are much Fig. 1b). The sum in (20) is replaced by a linear com- smaller in the prior solution than in the ADP data (cf. bination of the representers of the reduced basis [see Figs. 3b, and 3a), suggesting that most baroclinic signal (B1)]. A vector bˆ of 68 representer coef®cients is sought propagates into the study area from outside. It is thus to minimize the original cost function (19), that is, with essential to estimate baroclinic ¯uxes at the open bound- all the data in the second penalty term. Note that no ary (OB). This is done by means of DA. actual u and ␷ data are measured at the reduced basis To obtain an inverse solution, cov and covd should locations and the radial components from the two HF be speci®ed. A great deal of uncertainty still exists about radars do not have to be reprocessed to obtain u and ␷ error statistics of HF radar data. Contributing factors components. The inverse solution ␯Ã inv obtained with the include instrumental and processing errors, environ- reduced representer basis should be a close approxi- mental conditions that affect signal scattering, and rep- mation of ␯inv, obtained with the full basis, provided resentation of the measurement in the dynamical model representers for all the radial data can be approximated (i.e., the choice of lk). In the absence of detailed infor- well as a linear combination of representers from the mation about all these factors, in this study we assume reduced basis set. With this approach, the same rela- the simplest diagonal covariance matrix for data errors, 2 Ϫ1 tively small representer basis is used to ®nd the solution covd ϭ ␴ dI, where ␴d ϭ 0.02 m s is based on typical in every time window, while the number and location MLS errors of the harmonic constants, and I is the iden- of harmonically analyzed HF radar data meeting our tity matrix. accuracy threshold (see section 2) vary with time. 4. Error covariance for the OB condition d. Computational setup Although errors in the linearized equations are surely Most computations have been performed for the 38 not zero, we hypothesize that the major source of so- km ϫ 56 km area shown in Fig. 1b that has a maximum lution error in our small domain is the open boundary depth of 250 m and includes Stonewall Bank, a moderate condition (10) and take this to be the only weak con- bathymetric rise. In this area, the data from both HF straint. Thus, the rhs of Eqs. (11)±(14) will all be 0 radars, at Yaquina Head and Waldport, provide infor- except in the rows corresponding to open boundary mation about two components of surface currents (Kos- nodes, where fu is equal to the speci®ed barotropic cur- ro et al. 1997). The computational grid has 1-km res- rent plus the error that is estimated through DA. Ac- olution in the horizontal and 21 ␴ levels in the vertical. cordingly, in the matrix cov [see (21)] the only nonzero Some computations described in sections 4 and 5 have entries are the elements corresponding to the spatial also been performed for larger domains, at 2- and 3-km covariance of OB ¯uxes: C(xj, ␴j; xk, ␴k) ϭ͗q(xj, resolution. ␴j)qЈ(xk, ␴k)͘, where (x, ␴) ϭ (x, y, ␴) are positions on Some dissipation is needed for numerical stability, to the OB, and q represents u at the western boundary and smear sharp gradients across wave beams resulting from ␷ at the northern or southern boundary of our rectangular the singular forcing of the adjoint solution [see (21)]. domain. Vertical eddy viscosity is taken to be horizontally uniform In an oceanographic context, the input error covari- Ϫ2 4 and varying with depth as KM ϭ 10 exp[Ϫ(z/10) ] ϩ ances are known only approximately, at best. In practice, 1740 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 33 the choice of covariance is guided by both the need to OB condition. The Type II covariance for u and ␷ at satisfy (at least approximately) a number of physical the boundary of the local grid can be obtained as a result constraints and by ease of implementation. In our case, of the representer computation on a larger, coarser-res- physical constraints with which the open boundary co- olution grid. In our application, the local grid is nested variance must be consistent include the following: (i) in a large-scale grid covering an area of 177 km ϫ 435 at the OB, correction is provided only to deviations from km with 3-km resolution in the horizontal and 11 ␴ the depth-average current (recall that deviations from surfaces in the vertical (this domain is larger than the the depth-average at the OB are 0 in the prior solution, area shown in Fig. 1a); maximum water depth is set to while depth-averaged currents are believed to be rea- 500 m. For the large-scale grid, the open boundary con- sonably accurate); (ii) u at the west, and ␷ at the north dition is a strong constraint, and the horizontal mo- and south boundary segments are correlated in a way mentum equation (2) is a weak constraint. To form ma- to avoid nonphysical solution behavior in the corners trix cov for the large-scale model, the covariance for of the computational domain; (iii) the expected prior the errors in each u and ␷ component of Eq. (2) is written error of the velocity at the OB is of the order of a few as in (24) with xj and xk now inside the domain. Here centimeters per second (this requirement can be readily C1 is Gaussian with a decorrelation length scale of 10 satis®ed by scaling the covariance); (iv) low vertical km and C2 is as above, based on the local ¯at bottom modes are dominant; and (v) the covariance has a dy- modal decomposition. The Type II covariance is com- namically consistent spatial structure since, for instance, puted on the large-scale grid as a representer matrix for propagating modes are correlated at different parts of velocity observations at the locations corresponding to the boundary in accordance with their phase speed and the OB nodes of the small, nested domain (observations direction of propagation. This list of constraints for the are for u at the west, and ␷ at the north and south). covariance can be re®ned and completed as more ex- Figure 5 illustrates the difference in the spatial pattern perience is gained in regional DA modeling, for ex- of the Type I and Type II covariances at the northern ample, via the analysis of the representers and the in- boundary. Distinctive features of the Type II covariance verse solution. are determined by the superinertial wave dynamics in We tested two covariances that differ in the extent to the strati®ed medium. The ®rst baroclinic mode is pre- which criterion (v) is addressed. The Type I covariance served, indicated by the 180Њ phase difference between is constructed in an ad hoc manner guided by the prin- the surface and bottom. In contrast to the strictly real- ciples outlined above. The hope with such an approach, valued Type I covariance, the Type II covariance is which has been frequently used in the past, is that the complex-valued allowing for progressive phase changes covariance enforces smoothing and thus regularizes the along the boundary. The fact that energy (and hence inverse problem, but details of structure do not matter information) propagates along wave characteristics so much. The Type II covariance is obtained following manifests itself in the pattern of phase lines oriented at a nesting approach, using the representer solution in a an angle to the horizontal. For the assumed␳ (z), the larger domain. bottom slope near the coast is close to critical (see Fig. The Type I covariance is assumed to be separable as 1a), and here energy density may be intensi®ed near the bottom. Accordingly, the covariance amplitude, as well C(x , ␴ ; x , ␴ ) ϭ C (x , x )C (␴ , ␴ ), (24) jjkk 1 jk2 jk as the expected prior error, is increased at this spot. where C1 and C 2 are both symmetric and positive def- In the ``hand-made'' Type I covariance (see Fig. 5b), inite. To ensure regularity of u and ␷ around corners the ®rst baroclinic mode is also dominant, by construc- (criterion ii), velocity u in a ␴ layer is partitioned into tion. However, the other dynamically realistic features a nondivergent ¯ow, described by a streamfunction ␺, evident in the Type II covariance are missing. and an irrotational ¯ow, described by a potential ␾. Spatial correlations for both ͗␺␺͘ and ͗␾␾͘ are assumed 5. Computations with synthetic data: Effect of the to have a Gaussian form, with ͗␺␾͘ϭ0. Then C is 1 OB condition error covariance obtained in the standard way as a linear combination of derivatives of ͗␺␺͘ and ͗␾␾͘. To derive C 2, the errors When inverting HF radar surface currents (section 7), in the OB boundary condition are projected onto local ADP velocities are available to validate the data assim- ¯at bottom rigid-lid baroclinic vertical modes associated ilation results only at one location. To see better how with strati®cation N(z), and the modal amplitudes are the choice of the covariance for the open boundary con- assumed to be uncorrelated (Kurapov et al. 2002). dition (section 4) affects the solution everywhere in the Our construction of the Type I covariance is already computational domain, tests with synthetic data have quite involved, and yet some rather obvious issues (e.g., been performed. To generate synthetics we ®rst compute appropriate phase velocities along the boundaries) have the model solution for the larger 94 km ϫ 150 km area not even been considered. Nesting, based on the inter- shown in Fig. 1a, forced with depth-averaged currents pretation of the representers as covariances [see (23)], from the shallow-water inverse model. In this area, an provides a simpler and more general approach to con- internal tide with an amplitude comparable to obser- struction of a physically consistent covariance for the vations at the ADP site is generated over the shelf break. AUGUST 2003 KURAPOV ET AL. 1741

FIG. 6. Effect of the OB error covariance on the rms difference

between the inverse and validation u for a range of weights wo in the computations with synthetic data. Thick lines: total rms; thin lines: depth-average rms at the ADP location. Solid lines correspond to the Type II covariance (obtained by nesting) and dashed lines to the Type I covariance.

puts can be controlled by scaling the covariances in (19). Ϫ1 Let us replace the covariance cov with w o cov, where wo is a scaling factor for the open boundary penalty term in the cost functional (we again consider a strong constraint case, so only OB elements of cov are non- zero). The rms error of the inverse solutions, computed with the Type I and Type II covariances, is plotted as

a function of wo in Fig. 6, where the total rms error is averaged over the whole domain. For large values of

FIG. 5. Error covariance of ␷ at the location shown as the star and wo the solution rms error is close to that for the prior ␷Ј everywhere else at the northern boundary: (a) Type II covariance solution. The minimum rms error is attained for wo ϭ (obtained by nesting), complex amplitude is indicated by shading, 1 both for Type I and Type II covariances. For lower phase is contoured; (b) Type I covariance [real valued, 90Њ phase line w , the surface data is ®t better, but the DA model divides positive (at the top) and negative (near the bottom) values]. o performance deteriorates. Very low values of wo cor- respond to the case of effectively no open boundary This solution is obtained with resolution 2 km ϫ 2km condition penalty term in the cost function. The problem in the horizontal, and 11 ␴ surfaces in the vertical, with becomes ill-conditioned; solution irregularities originate the maximum water depth set at 500 m. The total, depth- near the open boundary and propagate inside the domain dependent currents obtained from the 2-km resolution (Foreman et al. 1980; Kivman 1997). In our case, spu- solution are then used to force the model in the small, rious small scale baroclinic vertical modes appear in the higher-resolution domain to generate a validation so- solution for low wo. lution. Synthetic u and ␷ data are sampled from this For the full range of wo, the inverse solution obtained solution at 34 surface locations, shown as circles in Fig. with the Type II covariance has a smaller total rms error 1b, and random noise of amplitude 0.02 m sϪ1 is added. than that obtained with the Type I covariance. Further-

The inverse solution obtained with these synthetic data more, for wo Ͻ 1 the Type II solution is much less is compared with the validation solution by computing sensitive to weight misspeci®cation than Type I. Over- the rms difference of the complex-valued three-dimen- estimating the expected open boundary error amplitude sional velocity ®elds. by an order of magnitude corresponds to wo ϭ 0.01 and Since the choice of covariances in the cost function for this weight the Type I covariance yields a solution (19) is always a hypothesis, it is important to learn about with rms error already worse than the prior, while the sensitivity of the inverse solution to their speci®cation. Type II solution still provides improvement. For small

In reality, uncertainty exists ®rst of all about the mag- wo, the total rms error is larger than the depth-average nitude of the errors, both for the model and the data. rms error at the ADP location, indicating that the so- The relative amplitude of the assumed errors in the in- lution quality is worse at the boundaries. This is illus- 1742 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 33

FIG. 7. Effect of the OB covariance on the spatial pattern of the depth-averaged rms error, in the computations with

synthetic data: (a) rms of the prior u; (b) rms of the inverse u, obtained with the Type II covariance, wo ϭ 1; (c) same Ϫ2 as (b), but with underspeci®ed weight wo ϭ 10 ; (d) rms of the inverse u, obtained with the Type I covariance; (e) same Ϫ2 as (d), but with underspeci®ed weight wo ϭ 10 . trated in Fig. 7 with 2D plots of depth-averaged rms. ditions are used, but treated as a weak constraint, a poor

Although for wo ϭ 1 the two covariances yield solutions choice of covariance can result in partial re¯ection. Con- of comparable quality (cf. Figs. 7b and 7d), for the Type versely, in data assimilation there is an opportunity to I covariance the solution corresponding to an under- make re¯ective boundary conditions more nearly radi- estimated weight (wo ϭ 0.01) is much more irregular ative by an appropriate choice of the covariance. We (cf. Figs. 7c and 7e). Thus, results from experiments hypothesize that the improved performance of the Type with synthetic data indicate improvements with the Type II covariance results at least in part from improved ra- II covariance, which we use to obtain results in the next diation. The Type II covariance should have better ra- sections unless otherwise speci®ed. diation properties since it is computed in a larger domain Note that the choice of the open boundary error co- that ``does not know'' about the existence of the small variance affects transparency of the boundary to out- domain. Improvement of wave radiation by the choice going waves. Even if perfect radiation boundary con- of the OB covariance in a DA model should be explored AUGUST 2003 KURAPOV ET AL. 1743

FIG. 8a. Component ␷ of the representer for the surface ␷ measurement computed with Type II covariance. (upper) South±north (SN) and west±east (WE) cross sections through the observation location (at 44.66ЊN, 235.7ЊE marked as the star). (lower) Plan view of representer on the surface. Here and in Figs. 14 and 15 amplitude is shaded, and phase angle, contoured, increases in the direction of phase propagation.

FIG. 8b. As in Fig. 8a but for observation at 44.58ЊN, 235.6ЊE over Stonewall Bank. 1744 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 33 in detail in future studies, since it may be applicable in coastal data assimilation in a broader context.

6. Representer structure Analysis of the representer is useful since, for in- stance, it shows how information from the data projects (in space and time) into the inverse solution. Figure 8 shows the ␷ component of the representer for a surface ␷ measurement. Zero phase contours in the alongshore, south±north (SN) section extend from the observation point downward along wave characteristics, giving a FIG. 9. Rms difference of the ADP validation data and model so- clear illustration of the direction of information prop- lutions, for each time window: prior solution (dashed thick line), agation. For a measurement made northeast of Stonewall inverse solution with all the radial data assimilated (thin dot±dashed Bank (Fig. 8a), two such zero phase lines are apparent line), inverse solution with radial data assimilated west of 235.8ЊE to the north and south of the observation point. In the (solid thick line), and inverse solution with data assimilated west of 235.7ЊE (solid thin line). west±east (WE) cross-shore section, the representer structure is similar to that of the Type II input covariance at the northern boundary (see Fig. 5a) with a compli- a. Comparison with ADP validation data cated phase pattern resulting from superposition of in- Figure 9 shows the depth-average rms difference be- coming and re¯ected waves. For a ␷ measurement made tween the computed solutions and ADP validation data over Stonewall Bank (Fig. 8b), the representer structure for each of the 16 overlapping 2-week time segments. is more strongly in¯uenced by bathymetry. Slopes of Assimilation of the whole set of HF radar harmonic phase contours are close to the bathymetric slope at the constants yields a reduction in rms mis®t compared with northern ¯ank of the bank; that is, the slopes are close the prior solution only for days 151±167 (compare the to critical. thin dot-dashed and thick dashed lines in Fig. 9). When In both Fig. 8a and 8b, the representer phase on the only data west of longitude 235.8Њ are assimilated, im- surface implies propagation from northwest to south- provement is obtained for a number of earlier and later east. This pattern may to some degree result from the time segments (thick solid line in Fig. 9). For days 171± OB covariance (Type II), and hence re¯ect dynamics in 179 the inverse solution rms mis®t is less than the prior the larger domain. However, a similar pattern can be solution rms mis®t if only radial currents west of lon- seen in the representer obtained with the Type I co- gitude 235.7Њ are assimilated (thin solid line in Fig. 9). variance, so the preferred phase direction may well be However, data in the stripe 235.7ЊϽlon Ͻ 235.8Њ still the effect of the local dynamics. contain valuable information about ¯ow in the begin- A representer for a sea surface elevation is qualita- ning of the assimilation period (compare the thin and tively similar to that for a surface velocity, suggesting thick solid lines for days 139±147). The reductions in that ␩ measurements, for instance, from a satellite al- mis®t that result from not ®tting the near-coast data timeter, may in principle be a valuable source of infor- suggest that the model assumption of horizontally uni- mation about baroclinic tides. However, the computed form background strati®cation may be poor for at least expected prior error for ␩ in our domain is less than 1 a part of the study period. Since the winds during the cm. A signal of such small amplitude contaminated by study period are generally upwelling favorable (see Fig. observational noise would be dif®cult to observe on the 4), isopycnals are presumably lifted near the coast (al- Oregon shelf. though we have no measurements to verify this). During short relaxation events (e.g., days 155±157 and 160± 7. Inversion of HF radar surface currents off 162; see Fig. 4) the pycnocline will not return to hor- Oregon, May±July 1998 izontal if the cross-shore pressure gradient is geostroph- ically balanced by alongshore currents, and so devia- Analysis of the representer suggests that information tions from our assumed, horizontally uniform strati®- from the surface velocity measurements propagates to cation will tend to increase throughout the study period. depth along wave characteristics. This fact, as well as Our model is likely to get worse as the upwelling con- results from the experiments with synthetic data, shows tinues. In these conditions, model wave characteristics that in principle assimilation of harmonically analyzed near the coast may especially be expected to have the surface currents can be useful for predicting internal wrong slope, so representers will project information tidal ¯ows at depth. Assimilation of real data allows a from the near-coastal data to the wrong locations off- rigorous test of the actual value of this data and our shore. model. It also shows where efforts should be ®rst di- The inverse solution does not offer improvement in rected to improve the inverse model. terms of rms for days 131 and 135 when the ADP data AUGUST 2003 KURAPOV ET AL. 1745

FIG. 10. Ellipses of horizontal baroclinic velocity at the ADP data locations for days 171±191, obtained by (a) assimilating HF radar radials only west of longitude 235.7ЊE (instead of 235.8ЊE as plotted in Fig. 3c), or (b) assuming stronger strati®cation (dashed line in Fig. 2) in the ®rst 50 m. Both of these variants on the standard case of Fig. 3c have surface-intensi®ed baroclinic currents, similar to the ADP data.

show low baroclinic signal (see Fig. 3a), or beyond day q12Јq CC(q , q ) ϵ |CC|ei␾CC ϭ (25) 179. Incorrect background strati®cation in the model 12 1/2 1/2 (qЈ11q )(q 22Јq ) may explain the poorer performance at the end of the studied period. Note that the relative increase in the rms and the gain G, which complements | CC | in quanti- fying the relative magnitude of q and q : mis®t after day 159 coincides with the rise in baroclinic 1 2 current amplitudes (see Fig. 3a). |qЈ12q | Figure 3c shows horizontal tidal ellipses for the com- G(q12, q ) ϭ . (26) qЈ22q puted baroclinic currents at the ADP location for the Here the vector of complex-valued entries q represents case lon(HF) Ͻ 235.8Њ. Qualitatively, the inverse model 1 the computed solution, q the corresponding validation captures much of the internal tide variability in the val- 2 data, and prime again denotes complex conjugate trans- idation data (cf. Fig. 3a). Starting with day 171, the pose. The vectors q and q may represent vertical pro- ADP data and inverse solutions have signi®cantly dif- 1 2 ®les of currents for a given day, or alternatively time ferent phases near the surface. Also, in the ADP data series of harmonic constants at a given depth. the internal tide is surface intensi®ed, a feature not re- Covariability in the vertical is assessed for each day produced by the inverse solution. Surface intensi®cation for the case lon(HF) Ͻ 235.8Њ (Fig. 11), separately for is reproduced in the inverse solution if more data are u (dashed lines) and ␷ (solid lines). The prior solution, dropped near the coast [case lon(HF) Ͻ 235.7Њ, Fig. which does not vary in time, is used to provide reference 10a]. Alternatively, surface-intensi®ed currents are also levels for all parameters (shown in Fig. 11 as thin lines). obtained if the assumed strati®cation is increased in the The inverse solution (bold lines) captures variability in upper 50 m (on a ¯at bottom, such change in strati®- the vertical better than the prior solution in terms of cation would have a similar effect on the shape of the each | CC |, G, and ␾CC for most of the study period. ®rst baroclinic vertical mode). Strati®cation correspond- Covariability with time is assessed at each vertical level ing to the latter experiment is shown as a dashed line where ADP data is available. When the whole time se- in Fig. 2, and the resulting baroclinic tidal ellipses at ries (days 131±191) is used, the inverse solution is of the ADP location in Fig. 10b. better quality than the prior solution in terms of | CC | Covariability of the validation data and computed so- and G in the larger part of the water column. Improve- lutions can be quanti®ed by the complex correlation ment in terms of ␾CC is marginal. If the statistics are 1746 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 33

Our results suggest that the assumption of horizon- tally uniform background strati®cation is a signi®cant de®ciency in our model. To the extent that this is true, our initial hypothesis that the only signi®cant source of error is in the open boundary forcing is questionable. The hypothesis about the errors can be tested formally (Bennett 2002, his section 2.3.3). Under the assumption that the errors are Gaussian with zero mean and co-

variance cov and covd, twice the cost function 2J(␯inv) is itself a random variable having a ␹ 2 distribution with 2K degrees of freedom (a factor of 2 appears since the data are complex valued); recall K is the number of assimilated data (K ഠ 540 in the case lon(HF) Ͻ 235.8Њ).

In particular, the mean value of 2J(␯inv) should be 2K and the variance 4K. Note that the reduced basis estimate

J(␯Ã inv) (B2) is larger then the full basis estimate J(␯inv),

although these values should be close since ␯Ã inv ഠ ␯inv.

For our series of inverse solutions J(␯Ã inv)/K is in the range of 2±12 implying that hypothesis about errors

should be rejected. The data term in J(␯Ã inv) accounts for about 95% of the total cost function value, so changing

covd should have a stronger effect on J(␯Ã inv) than cov. FIG. 11. Functions showing ADP-prior and ADP-inverse solution In this respect, better knowledge of the error model for covariability in the vertical, for each day: amplitude | CC | and phase the data would be desirable. ␾CC of complex correlation, and gain G (26), where q2 represents the ADP measurements; lon(HF) Ͻ 235.8ЊE. By itself the ␹ 2 criterion does not say where the de- ®ciency in our hypothesis is; e.g., either one or both of

cov and covd could be misspeci®ed. Increasing our es- computed only for days 139±167 for which the solution timate of the data error standard deviations from 2 to 3 Ϫ1 rms error is improved, the inverse solution is of better cm s decreases J(␯Ã inv)/K to a value near 1 over much quality than the prior solution in terms of all the three of the ®rst half of the study. Such an increase in data criteria (Fig. 12). Note that at middepths, where the error is quite plausible. However, in the second half of magnitude of the baroclinic currents is low, estimates the study, when J(␯Ã inv)/K Ͼ 6, unreasonably large in- of CC and G lose meaning. The fact that all the criteria creases in the data error would be required to bring the are improved below 40 m provides a clear demonstration test statistics down to acceptable levels, implying that of the value of surface currents from HF radars in con- some modi®cations to cov are required at least for this straining subsurface tidal ¯ows. time period. Increasing the expected magnitude of errors

FIG. 12. Functions showing ADP-prior and ADP-inverse solution covariability with time, at each depth [lon(HF) Ͻ 235.8ЊE]: statistics for days 139±167, for which the solution rms error is improved. Legend as in Fig. 11. Values corresponding to rms amplitudes of the ADP ¯uctuation currents Ͻ 1cmsϪ1 are not shown. AUGUST 2003 KURAPOV ET AL. 1747

FIG. 13. Tidal ellipses of the inverse solution, day 139: (a) depth-averaged current and (b) deviations from the depth-average near the surface. Contours are isobaths of 50, 75, and 100 m. in the baroclinic boundary conditions has little effect mate. The horizontal tidal velocity vectors of the bar- on the value of J(␯Ã inv). Furthermore, similar values are otropic ¯ow rotate counterclockwise and ellipses are obtained for the test statistics with both the Type I and strongly polarized, consistent with shelf-modi®ed Type II covariances. This suggests that as long as we dynamics (e.g., Munk et al. 1970). At the retain a strong constraint formulation, we cannot make same time, vectors of the horizontal baroclinic current reasonable adjustment to cov and covd that are statis- (de®ned as the deviation from the depth-averaged) are tically consistent with observations. Only by admitting less polarized and rotate clockwise, consistent with ki- errors in our equations (and hence signi®cantly changing nematics of free planar waves. Indeed, the direction of our assumption about cov) could we achieve consis- rotation depends on the phase angle ␣ between complex tency between the HF radar data and our simpli®ed amplitudes u and ␷, such that, by our convention, sin dynamical model. ␣ Ͻ 0 corresponds to clockwise rotation. If the rotary Although our analysis of the ␹ 2 criterion suggests that components of the current are de®ned as ␷Ϯ ϭ (u Ϯ the inverse solutions satisfying our simpli®ed dynamical i␷)/2 then | ␷Ϯ | 2 ϭ (|u | 2 ϩ | ␷ | 2 Ϯ 2|u࿣␷ | sin␣)/4. equations should be considered suboptimal, they are still So, for the vector rotating clockwise, | ␷ϩ | Ͻ | ␷ Ϫ |. It a signi®cant improvement over the prior. Moreover, our is easily shown that for a planar wave (i.e., a wave simpli®ed model has allowed us to achieve the goals of involving a pressure gradient only in one direction) the study, stated in the introduction. From here we can | ␷ϩ |/|␷ Ϫ | ഠ (␻ Ϫ f )/(␻ ϩ f ), and so in our case improve our model or attempt to use the dynamics as | ␷ϩ | K | ␷ Ϫ | and rotation should be clockwise. To see a weak constraint. clearly the phase propagation of the baroclinic wave, we plot the ␷ Ϫ component of the baroclinic current at the surface for days 139, 155, and 167 (Fig. 14). A b. Variability of the tidal ®elds complicated phase pattern is seen and is due to the su-

The internal tide is a signi®cant contributor to M2 perposition of incident waves and those re¯ected from tidal ¯ow at the surface. Deviations at the surface fre- topography (or the coast). Despite day-to-day variability quently exceed the depth-averaged tidal current (Fig. in details, dominance of phase propagation from the 13). The computed internal tide varies on a scale of northwest to the southeast of the area is a persistent about 10 km, in accordance with the theoretical esti- feature. 1748 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 33

FIG. 14. Rotary current (u Ϫ i␷)/2 at the surface, deviations from the depth average, for days 139, 155, and 167. The star shows an ADP location.

The same phase propagation pattern can be found in (e.g., Gill 1982, chapter 10). This can be shown using isopycnal displacements. The dominance of shoreward an analytical model of the strati®ed ¯ow over the ¯at phase propagation is seen most clearly in the cross-shore bottom (Wunsch 1975), where the ¯ow is separated into section through the northern part of the study area, a series of vertical modes and the dynamics of individual where topography is relatively simple (Fig. 15, left pan- modes are governed by shallow-water-type equations: els). In the cross section through Stonewall Bank (right U f U P, (27) panels) topography complicates the phase pattern. For i␻ ϩ ϫ ϭϪ١ (U ϭ 0. (28 ´ ١(instance, for day 155, the phase of the isopycnal dis- i␻P ϩ (1/␭2 placements suggests scattering off Stonewall Bank to the west and east. The magnitude of the computed is- Here U(x, y) and P(x, y) are modal amplitudes of u and p/ , correspondingly. In the case N const, /(NH) opycnal displacements at middepth is 2±6 m, sometimes ␳o ϭ ␭ ഠ ␲ reaching 10 m over Stonewall Bank. for the ®rst baroclinic mode. If the coast is straight and The magnitude of the internal tide can also be quan- is aligned with the y axis at x ϭ 0, then the solution is sought as a superposition of an incident wave P P ti®ed by computing baroclinic kinetic energy (per unit ϭ a exp(Ϫikx Ϫ ily) and a wave re¯ected from the coast P ϭ mass) averaged over the tidal period, KEBC ϭ | u Ϫ 2 Pb exp(ikx Ϫ ily). Let us put Pa ϭ 1. To satisfy a no-¯ow u1 | /4, where u1 is the complex amplitude of the depth- boundary condition at x ϭ 0, Pb ϭϪc1/c1* , where c1 ϭ averaged current. Near the surface, superposition of 2 2 waves incident and re¯ected from the coast results in Ϫ(lf ϩ i␻k)/(␻ Ϫ f ), and the asterisk denotes complex conjugate. For a Poincare wave, k2 ϩ l2 ϭ ␭2(␻2 Ϫ f 2), zones of higher and lower KEBC.IfKEBC is averaged over a series of days, zones of higher internal tide energy and so the cross-shore component of the wave vector k near the surface appear to be aligned along the coast depends on the strati®cation, depth, and the direction ␣ (Fig. 16a). This pattern may super®cially suggest sur- of the incident wave phase propagation. Proceeding with face bounce points of beams of internal tide oriented the solution, one obtains 22 perpendicular to the coast. However, this pattern and |U| ϭ 2|c1|[1Ϫ cos(2kx)], (29) cross-shore spatial scales are in fact consistent with en- 22 ergetics of a Poincare wave in the presence of a coast |V | ϭ 2|c2|[1Ϫ cos(2kx ϩ ␾)], (30) AUGUST 2003 KURAPOV ET AL. 1749

FIG. 15. Amplitude and phase of isopycnal displacements, shown in two west±east (WE) cross sections, (left) one north of Stonewall Bank and (right) another through the bank, for days 139, 155, and 167. where U and V are cross-shore and alongshore compo- Near the bottom, the intensity of the internal tide is 2 2 nents of U, c2 ϭ (l␻ ϩ ifk)/(␻ Ϫ f ), and ␾ ϭ determined more directly by the interaction with local arg[c1c*21/(c*c2)]. Thus, near the coast, the kinetic energy topography, with larger KEBC found on the ¯anks of of a baroclinic mode [the sum of (29) and (30)] does not Stonewall Bank (Fig. 16b). In the vertical, west±east depend on the alongshore coordinate and has zones of cross section north of the bank (Fig. 16c), an energy higher and lower magnitude in the cross-shore direction. minimum at middepth indicates dominance of the ®rst Relevant to our study, we take an estimate of N ϭ 0.01 baroclinic mode. In the west±east section through Stone- Ϫ1 s , H ϭ 100 m, and ␣ ϭϪ45Њ (a wave from the north- wall Bank, high KEBC in the water column over the bank west) so that for baroclinic mode 1 | c1 | ϭ | c2 |, ␾ ϭ indicates that internal tide beams generated over this Ϫ34Њ, and ␲kϪ1 ϭ 15 km. Then, the ®rst three zones of topographic feature extend to the sea surface. max KEBC should be 7, 22, and 37 km from the shore, Day-to-day evolution of KEBC averaged over the tidal in close agreement with results of Fig. 16a. period and over the whole computational domain (Fig. 1750 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 33

FIG. 16. Plot of KEBC averaged over a series of days 139±167: (a) at the surface [contour lines are bathymetry every

20 m; red marks show locations of zones of maximum KEBC predicted by the theoretical Poincare wave solution (29)± (30)]; (b) at the bottom; (c) west±east cross section north of Stonewall Bank; and (d) west±east cross section through Stonewall Bank.

17) shows larger values between days 171 and 187, consistent with the impression from Fig. 3. Note that

the level of KEBC for days 139±147 is the same as for days 159±167 while both the validation data (see Fig. 3a) and the inverse solution at the ADP location (see Fig. 3c) show a lower baroclinic signal for days 139± 147 than for days 159±167. This suggests that for days 139±147 the baroclinic tide was stronger away from the ADP location than near the ADP site, as seen in Fig. 14a. So it may be misleading to attempt to assess the intensity of the internal tide in the whole area from information collected at one location.

c. Energy balance

FIG. 17. Plot of KEBC averaged over the whole computational Unless the bottom is ¯at, separation of a ¯ow into domain and over the tidal period, shown for a series of days. barotropic and baroclinic parts is not obvious, and the AUGUST 2003 KURAPOV ET AL. 1751

analysis and interpretation of energetics will depend to pressure, p 2(x, y, z, t) ϭ p Ϫ p1, and the ®rst term on some extent on how this partitioning is formally de®ned. the rhs of (31) is minus the barotropic energy ¯ux di-

If the barotropic current u1 is de®ned as the depth av- vergence; EC is the barotropic-to-baroclinic energy con- erage (Cummins and Oey 1997; Holloway 2001), and version rate: the baroclinic current as the deviation from the depth 0 average, u ϭ u Ϫ u , then the barotropic (depth in- cart 2 1 EC ϭ gw͵ Ã ␳ dz, (32) tegrated) energy equation is ϪH

cart -where wà is the vertical velocity (aligned with the ver 1 ץ 22 ␳o(g␩ ϩ H|u1|) tical z axis in Cartesian coordinates) associated with the :t 2 depth-averaged ¯owץ

␩ץ (Hu11p ) Ϫ ␳oru1 ´ u(ϪH) Ϫ EC. (31) z (z ϩ H)´ ϭϪ١ cart (١H ϩ . (33 ´ wà ϭ u1 tץHere the ®eld variables are time-dependent, rather than HH complex amplitudes, p1(x, y, t) is the depth-averaged The baroclinic, depth-integrated energy equation is

2 uץ00g␳2 0 1 ץ 2 (u22pdzϪ ␳oru2 ´ u(ϪH) Ϫ ␳o KdzM ϩ EC, (34 ´ ␳o |u2| ϩ dz ϭϪ١ zץ ͵ ͵ t 2 ͵ Ϫ␳ zץ ϪH ΂΃ϪH ϪH ΗΗ

where the ®rst term on the rhs is minus the baroclinic the shallowness of the area. Also, while the continental energy ¯ux divergence. De®ning the barotropic and bar- shelf slope outside our domain is steep, we do not expect oclinic ¯ows as above has the advantage that the energy much conversion here compared to Hawaii since bar- conversion term in (31) ®nds its exact counterpart with otropic ¯ow near the coast mostly follows bathymetric opposite sign in (34); adding (31) and (34) yields the contours (Baines 1982). energy equation for the total ¯ow so that the total energy If (33) is substituted into (32), the energy conversion in the model can be accounted for. term can be separated into two parts: In our analysis, we consider terms in the energy equa- ␩ g 0ץ (١H ϩ (z ϩ H)␳ dz, (35 ´ tions averaged over a tidal period. The depth-integrated EC ϭϪp | u ͵tHץ zϭϪH 1 2 barotropic energy ¯ux is from the south to the north ϪH (consistent with Kelvin wave dynamics), on the order where the ®rst term is the topographic energy conver- Ϫ1 of 10 kW m , with relatively smaller values over the sion (e.g., see Llewellyn-Smith and Young 2001). The shallows of Stonewall Bank (Fig. 18a). The barotropic second term is associated with sea surface variations, ¯ux has low day-to-day variability. The depth-integrated and its physical interpretation is not entirely intuitive. baroclinic energy ¯ux is much smaller, on the order of Its appearance is the consequence of our approximate tens of watts per meter, and experiences substantial tem- division into barotropic (depth averaged) and baroclinic poral variability (Figs. 18b±d). The persistent feature ¯ows, with the elevation and density variations attri- for all days studied is the general direction of baroclinic buted solely to the barotropic and baroclinic components energy ¯ux from northwest to southeast, close to the respectively. Consideration of the classical problem of direction of phase propagation. This suggests that the internal waves over a ¯at bottom with a free surface continental shelf break slope northwest of the studied (Phillips 1966; Wunsch 1975) shows that this division area may be a generation site of energetic M 2 internal is not exactly correct. For that simple analytical case tide which can propagate onto the shelf. Internal tides the ¯ow can be decomposed into a series of vertical are probably also generated due west and south of our modes such that the depth averages of the baroclinic computational domain, but the steeper slopes in these pressure modes are not identically zero. Potential energy areas (Fig. 1a) re¯ect most of the internal tide energy for both the barotropic (mode 0) and baroclinic (all other toward deeper water. modes) components include contributions from both el- The time-averaged terms in the energy equations for evation and density variations. Since the modes are dy- day 139 are shown in Figs. 19a±d and 19g. The diver- namically uncoupled [cf. (27)±(28)] there will be no gence of the barotropic and baroclinic energy ¯uxes, barotropic to baroclinic energy conversion. In particular, and the energy conversion rate, are only a few milliwatts the second term in (35), which is present even over a per meter squared, at most. This is several orders of ¯at bottom in our approximate treatment, would not magnitude lower than the values seen in areas of strong appear with a proper division into barotropic and bar- internal tide generation such as the Hawaiian Ridge oclinic modes. (Merri®eld and Holloway 2002). This is partly due to In regions of strong bathymetric variation (such as 1752 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 33

FIG. 18. Energy ¯ux (EF), depth-integrated and averaged over the tidal period: (a) barotropic EF, for day 139, scale vector in the lower right corner is 15 kW mϪ1; (b)±(d) baroclinic EF, for days 139, 155, and 167, scale vector is 40 W mϪ1. Shaded contours are bathymetry, every 20 m.

ocean ridges or the continental shelf break) the ®rst positive conversion occurs at the north slope. Reverse term dominates in (35). However, within our study topographic energy conversion, from the baroclinic area topographic variation is relatively mild such that to barotropic ¯ow, is possible because the sign of the the vertical velocity at the surface is comparable with time-averaged EC depends on the phase shift between cart that at the bottom. Accordingly, both components of wà and p 2 at the bottom; since most of the baroclinic EC are of similar magnitude (cf. Figs. 19d±f and 19g± wave propagates into the domain from outside, p 2 is i, where the terms are shown for days 139, 155, and not necessarily phase locked to wà cart . 167). However, they have different, distinctive pat- The phase difference between ␳ and ␩ determines the terns. For days 155 and 167, topographic conversion sign of the second term of (35), averaged over a tidal is positive and strongest near the western slope of period. The plots for this term (see Figs. 19g±i) highlight Stonewall Bank. However, for day 139, the topo- the baroclinic wave propagating from the northwest and graphic EC is negative in the same area, and strong apparently re¯ecting from topography and the coast. AUGUST 2003 KURAPOV ET AL. 1753

FIG. 19. Terms in the energy equations averaged over the tidal period: (a) divergence of the barotropic energy ¯ux, day 139; (b) divergence of the baroclinic energy ¯ux, day 139; (c) total dissipation, day 139; (d)±(f) topographic EC for days 139, 155, and 167, respectively; (g)±(i) second term in (35), for the same days. Contours are bathymetry every 20 m.

Terms on the rhs of (34) averaged over the whole residual (numerical error) in the energy balance com- computational area and over the tidal period are shown putation. The energy conversion terms are much smaller in Fig. 20a, for every time window. Plotted are the than the OB EF, demonstrating that most baroclinic en- baroclinic energy ¯ux into the area through the open ergy comes through the open boundary with the net boundary [OB EF, the spatial integral of the ®rst term in¯ux balanced by dissipation. on the rhs of (34)], the dissipation [combining the sec- Integrated over the 13 km ϫ 22 km area of the Stone- ond and the third terms on the rhs of (34)], the energy wall Bank, the time-averaged energy balance (Fig. 20b) conversion split into the two terms [see (35)], and the differs from that integrated over the whole computa- 1754 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 33

Oregon shelf. Comparison with the ADP data shows that assimilation of harmonically analyzed surface current measurements from the coast-based HF radars captures temporal variability, as well as variability in the vertical, of the internal tide.

Although M 2 baroclinic ¯ows on the mid±Oregon shelf are variable and intermittent, some features are found to be persistent. For instance, during the study period the phase and energy of the baroclinic tide prop- agated in the same direction, from north-west to south- east. At the surface, baroclinic surface tidal currents (deviations from the depth-averaged current) can be 10 cm sϪ1 , 2 times as large as the depth-averaged current. Allowing for spring±neap variations and more rapid changes in the internal tide in response to changes in ocean conditions, peak baroclinic velocities in the semidiurnal band could be easily 2 times this. The analysis of the energy balance demonstrates that most baroclinic signal comes into the study area via the open boundaries, and so a proper speci®cation of OB con- ditions is critical. The generalized inverse method offers great ¯exi- bility in the formulation of the DA problem. For in- stance, weak constraint DA (where dynamical equa- tions are assumed to contain error) is used here to ®nd an appropriate covariance for the open boundary con- dition by nesting. Then, strong constraint DA (with exact dynamics) is implemented to estimate open boundary baroclinic currents using the measured sur- face velocities. One of the values of the representer method is that a dynamically consistent prior error co- FIG. 20. Terms in the baroclinic energy equation (34) averaged over the tidal period and (a) averaged over the whole computational variance for the model solution can be computed, rather domain, or (b) averaged over the area of the Stonewall Bank. Case than guessed. lon(HF) Ͻ 235.8ЊE. Dots show the residual (numerical error) in the From the perspective of data assimilation, this study energy balance. raises important questions concerning open boundary conditions. Since DA can provide corrections to the tional domain (Fig. 20a). Over the bathymetric rise, boundary values, there is an opportunity to control topographic EC is comparable to the energy ¯ux into radiation by the choice of open boundary condition the smaller area and is biased toward positive values. covariance, consistent with the modeled dynamics. In our study, nesting is adopted as the simplest way to obtain a covariance that should improve wave radia- 8. Summary tion. Computations with synthetic data show that such Our study is the ®rst attempt to model superinertial a covariance yields an inverse solution superior to that internal tide variability on the shelf using data assim- obtained with the best covariance that we could make ilation. Since the tidal ¯ow depends critically on hy- up without nesting. drography and open boundary currents that are poorly Although the dynamical model can in principle be known, data assimilation seems a promising ap- used at any frequency, limited accuracy of HF radar proach. The representer solution and computations data poses dif®culties in estimating major tidal con- with synthetic data show that in principle measure- stituents, other than M 2 , from short time series. Off ments of surface currents can be used to map tidal Oregon, S 2 is the second largest tidal constituent with

¯ow at depth. For the superinertial M 2 tide, infor- barotropic tidal magnitude about 40% of M 2 and a mation from the surface is projected in space along period (12 h) very close to that of M 2 (12.4 h). The wave characteristics. Application of our frequency K1 diurnal tidal frequency is subinertial so that free domain inverse model to HF radar data harmonically internal waves are not allowed. Tidal signal in HF radar analyzed in a series of overlapping 2-week windows surface data at the K1 frequency is probably also con- provides a uniquely detailed picture of the temporal taminated by wind-induced currents due to a diurnal and spatial variability of internal tide on the central sea breeze (Erofeeva et al. 2003). Tidal studies in the AUGUST 2003 KURAPOV ET AL. 1755 diurnal band should probably make allowance for this small size matrix inversions locally in each vertical additional forcing. column of the computational grid. The internal tide inverse solution is sensitive to the To accomplish this, we approximate (11) by background strati®cation and velocity ®elds. These u ϭ Cf Ϫ ⌿ ␳ϩ , (A1) could possibly be accounted for within the context of uu our simple model by means of weak constraint DA. where ⌿ u ϭ C(G | P), Then careful interpretation of the dynamical error ␩ would be necessary to avoid confusion in the analysis ␳ϩ ϭ , (A2) of the momentum and energy budget. A better ap- ΂΃␳ proach to improvement of the internal tide prediction and C ഠ ⍀ Ϫ1 . For the original continuous equations would be to improve the model by allowing for a the operator corresponding to ⍀ in (11) and its inverse horizontally nonuniform background strati®cation. are local in the horizontal coordinates x and y. So, a Such modi®cations to the model should also include continuous analog of (A1) is exactly equivalent to (2). the background current to geostrophically balance However, discretized on the staggered C grid the Cor- background horizontal pressure gradients. Lineari- iolis term couples the u and ␷ components of u at zation with respect to this new basic state would yield neighboring horizontal locations. This destroys the dynamical terms that describe advection of tidal per- block structure of ⍀ and makes ⍀ Ϫ1 a full matrix. turbations by the background currents. Such a model To avoid this we directly discretized the horizontally could be useful for addressing the question of the local inverse operator: {u, ␷} is ®rst transformed into effect of subinertial currents on internal tides, which ␷ Ϯ ϭ (u Ϯ i␷)/2 (Lynch et al. 1992; Muccino et al. remains open. However, as soon as advective terms 1997); the momentum equations are decoupled with are included, the task of ®nding OB conditions that respect to ␷ ϩ and ␷ Ϫ , so inversion for ␷ Ϯ is performed guarantee a well-posed formulation becomes nontriv- locally at each u and ␷ location of the C grid; ®nally, ial (see Bennett 1992, chapter 9). Our example, with backward transformation is made to get u. All these a simpler model, shows that ®nding an appropriate steps are ®xed in the matrix C. set of OB conditions does not solve all the problems From (13), we obtain, taking (A1) into account, at the OB if the task is to restore OB inputs by data w WϪ1 f WDCϪ1 f WDϪ1 ϩ. (A3) assimilation. Attention should thus also be given to ϭ wuuϪ ϩ ⌿ ␳ the OB error covariance. The resulting system of equations for ␳ϩ is obtained by To provide a more realistic background state, a substitution of (A1) and (A3) into (12) and (14) and coastal DA system should eventually couple a tidal then combining them to give inverse model and a reliable model of wind-driven A␳ϩ ϭ F f ϩ F f ϩ f , (A4) circulation. This would provide a more realistic back- uwuw␳ϩ ground state for the tidal model, as well as a frame- where work for studying tidal effects on subtidal ¯ows. Our i␻I 0 success with inversion of the HF radar data to describe A ϭ B22ϩ D ⌿u, B 2ϭ , three-dimensional superinertial tidal ¯ow, using sim- ΂΃0 Η B ple dynamics, is encouraging for the optimal use of D surface current measurements with more advanced D ϭϪ a , 2 D ZWϪ1 D models. ΂΃␳␳Ϫ 0 Acknowledgments. The research was supported by Fu ϭ DC2 , Fw ϭ , ϪZWϪ1 the Of®ce of Naval Research (ONR) Ocean Modeling ΂΃␳ and Prediction Program under Grant N00014-98-1- fa 0043, and the National Science Foundation under f␳ϩ ϭ . f Grant OCE-9819518. ΂΃␳

The matrix B2 is partitioned into blocks such that stan- APPENDIX A dard block-matrix multiplication can be performed for ϩ B2 and ␳ (A2); Fw has zeros in the rows corresponding Model Solver to the depth-integrated continuity equation. The solution to (A4) is found by direct LU factorization of A using The solution strategy for (11)±(14) is to operate a standard routine for banded matrices (Anderson et al. with parts of the discretized model operator as sparse 1992). Then multiplication by both AϪ1 and its complex matrices, eliminate u and w, then solve the reduced conjugate, matrix transpose (AЈ)Ϫ1 can be accomplished size matrix problem by direct factorization. Elimi- rapidly by solving the appropriate triangular systems. nation of variables is accomplished by a series of Last, the solution to (11)±(14) is 1756 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 33

Ϫ1 Ϫ1 Ϫ1 C Ϫ ⌿uuuAFuwϪ⌿ AF Ϫ⌿ A uf      u  ␯ ϵ w ϭ ⌿ AFϪ1 Ϫ WDCϪ1 ⌿ AFϪ1 ϩ WϪ1 ⌿ AϪ1 f , (A5)   wwwwuw  ϩ ␳   f␳ϩ Ϫ1 Ϫ1 Ϫ1 AFuw AF A

Ϫ1 ÐÐ, 2002: Inverse Modeling of the Ocean and Atmosphere. Cam- where ⌿w ϭ W D⌿u, and a 3 ϫ 3 block-matrix is SϪ1. The size of the numerical problem is limited by bridge University Press, 234 pp. Blumberg, A. F., and G. L. Mellor, 1987: A description of a three- the storage required for the LU factors of the banded dimensional coastal ocean circulation model. Three-Dimensional matrix A. The advantage of the solution approach de- Coastal Ocean Models, N. Heaps, Ed., Coastal and Estuarine scribed above is that once matrix A is factored, a large Science Series, Vol. 4, Amer. Geophys. Union, 1±16. number of representers [see (21)] can be rapidly com- Bogden, P. S., 2001: The impact of model±error correlation on re- puted as a series of matrix-vector and vector±vector gional data assimilative models and their observational arrays. Ϫ1 J. Mar. Res., 59, 831±857. operations. Also, the adjoint solver (SЈ) is easily ob- Cummins, P. F., and L. Y. Oey, 1997: Simulation of barotropic and tained as a complex conjugate, matrix transpose of the baroclinic tides off northern British Columbia. J. Phys. Ocean- block matrix in (A5). ogr., 27, 762±781. Egbert, G. D., and S. Y. Erofeeva, 2002: Ef®cient inverse modeling APPENDIX B of barotropic ocean tides. J. Atmos. Oceanic Technol., 19, 183± 204. ÐÐ, A. F. Bennett, and M. G. G. Foreman, 1994: TOPEX/POSEI- Reduced Basis Approach DON tides estimated using a global inverse model. J. Geophys. LetlˆЈ be a data functional corresponding to the assumed Res., 99 (C12), 24 821±24 852. k Erofeeva, S. Y., G. D. Egbert, and P. M. Kosro, 2003: Tidal currents observation from the reduced basis, k ϭ 1, ´´´, K1.In on the central Oregon shelf: Models, data, and assimilation. J. our study lˆЈku sample u and ␷ at 34 points on the surface, Geophys. Res., 108, 3148, doi:10.1029/2002JC001615. so K1 ϭ 68. Representers of the reduced basis are com- Foreman, M. G. G., L. M. Delves, I. Barrodale, and R. F. Henry, Ϫ1 Ϫ1 1980: On the use of the Proudman±Heaps tidal theorem. Geo- puted as rÃk ϭ S cov(SЈ) ˆlk. The inverse solution is sought in the form phys. J. Roy. Astron. Soc., 63, 467±478. Gill, A. E., 1982: Atmosphere±. Academic Press, K1 662 pp. ␯à ϭ ␯ ϩ bˆ rà . (B1) Hayes, S. P., and D. Halpern, 1976: Observations of internal waves inv 0 ͸ kk kϭ1 and coastal upwelling off the Oregon coast. J. Mar. Res., 34, 247±267. The task is to ®nd the vector bˆ of representer coef®cients Holloway, P. E., 2001: A regional model of the semi-diurnal internal bˆ k that yields a minimum of the original cost function tide on the Australian North West Shelf. J. Geophys. Res., 106, (19), which is 19 625±19 638. à Kivman, G. A., 1997: Weak constraint data assimilation for tides in J(␯à inv) ϭ bˆЈRbˆ ϩ (d Ϫ LЈ␯ 0 Ϫ Pbˆ)Ј the Arctic Ocean. Progress in Oceanography, Vol. 40, Perga- mon, 179±196. Ϫ1 ϫ covd (d Ϫ LЈ␯0 Ϫ Pbˆ). (B2) Kosro, P. M., J. A. Barth, and P. T. Strub, 1997: The coastal jet: Here Rà is the representer matrix of the reduced basis, observations of surface currents over the Oregon continental à shelf from HF radar. Oceanography, 10, 53±56. of size K1 ϫ K1, with elements Rkn ϭ lˆЈkrà n; P is of size Kurapov, A. L., G. D. Egbert, R. N. Miller, and J. S. Allen, 2002: K ϫ K1 with elements Pkn ϭ lЈkkrà n, wherelЈ are the Data assimilation in a baroclinic coastal ocean model: ensemble discrete data functionals for actual observations [see statistics and comparison of methods. Mon. Wea. Rev., 130, 1009±1025. (17)], and rà n is the representer from the reduced basis. Optimal representer coef®cients are found as in Egbert Llewellyn Smith, S. G., and E. R. Young, 2001: Conversion of the barotropic tide. J. Phys. Oceanogr., 32, 1554±1566. and Erofeeva (2002): Lynch, D. R., F. E. Werner, D. A. Greenberg, and J. W. Loder, 1992: bˆ ϭ RQà Ϫ1/2⌳(l ϩ ⌳ 2)Ϫ1VЈcovϪ1/2(d Ϫ LЈ␯ ), (B3) Diagnostic model for baroclinic, wind-driven and tidal circula- d 0 tion in shallow . Cont. Shelf Res., 12, 37±64. where the matrices V, ⌳ (diagonal), and Q are the Mellor, G. L., 1998: Users' guide for a three-dimensional, primitive result of singular value decomposition: V⌳QЈϭ equation, numerical ocean model. Program in Atmospheric and s.v.d.{cov Ϫ1/2PRà Ϫ1/2 }. Oceanic Sciences, Princeton University, Princeton, NJ. [Avail- d able online at http://www.aos.princeton.edu/WWWPUBLIC/ htdocs.pom/.] REFERENCES Merri®eld, M. A., and P. E. Holloway, 2002: Model estimates of M2 internal tide energetics at the Hawaiian Ridge. J. Geophys. Res., Anderson, E., 1992: LAPACK User's Guide. Philadelphia, 235 pp. 107, 3179, doi:10.1029/2001JC000996. Baines, P. G., 1982: On internal tide generation models. Deep-Sea Muccino, J. C., W. G. Gray, and M. G. G. Foreman, 1997: Calculation Res., 29, 307±338. of vertical velocity in three-dimensional, shallow water equation, Bennett, A. F., 1992: Inverse Methods in Physical Oceanography. ®nite element models. Int. J. Numer. Meth. Fluids, 25, 779±802. Cambridge University Press, 346 pp. Munk, W., F. Snodgrass, and M. Wimbush, 1970: Tides off shore: AUGUST 2003 KURAPOV ET AL. 1757

Transition from California coastal to deep-sea waters. Geophys. Sirkes, Z., and E. Tziperman, 1997: Finite difference of adjoint Fluid Dyn., 1, 161±235. or adjoint of ®nite difference? Mon. Wea. Rev., 125, 3373± Oke, P. R., J. S. Allen, R. N. Miller, G. D. Egbert, and P. M. Kosro, 3378. 2002: Assimilation of surface velocity data into a primitive equa- Smith, R. L., A. Huyer, and J. Fleischbein, 2000: The coastal ocean tion coastal ocean model. J. Geophys. Res., 107 (C9), 3122, doi: off Oregon from 1961 to 2000: Is there evidence of climate 10.1029/2000JC000511. change or only Los Ninos? Progress in Oceanography, Vol. 49, Oliger, J., and A. SundstroÈm, 1978: Theoretical and practical aspects Pergamon, 63±93. of some initial boundary value problems in ¯uid dynamics. SIAM Torgrimson, G. M., and B. M. Hickey, 1979: Barotropic and baro- J. Appl. Math., 35, 419±446. clinic tides over the continental slope and shelf off Oregon. J. Pacanowski, P. C., and S. C. H. Philander, 1981: Parameterization of vertical mixing in numerical models of tropical . J. Phys. Phys. Oceanogr., 9, 945±961. Oceanogr., 21, 1186±1201. Wijesekera, H. W., J. S. Allen, and P. A. Newberger, 2003: Modeling Petruncio, E. T., L. K. Rosenfeld, and J. D. Paduan, 1998: Obser- study of turbulent mixing over the continental shelf: Comparison vations of the internal tide in Monterey Canyon. J. Phys. Ocean- of turbulent closure schemes. J. Geophys. Res., 108, 3103, doi: ogr., 28, 1873±1903. 10.1029/2001JC001234. Phillips, O. M., 1996: Dynamics of the Upper Ocean. Cambridge Wunsch, C. H., 1975: Internal tides in the ocean. Rev. Geophys. Space University Press, 261 pp. Phys., 13, 167±182.