1. Introduction
The nature of the oceans, as masses of fluid interacting with solid boundaries, makes possible, as a particular type of ocean–coast interaction, the phenomenon of an oceanic jet impinging on a coastal boundary. Among the many consequences produced by such an interaction, two features are of principal importance for the jet and its surroundings. These are its evident deflection by the coastal boundary and its division into two branches flowing along the coast. In the steady state, which is the case considered in this work, the phenomenon is observed only in specific oceanic regions having suitable boundary conditions, for instance at the exits of straits where a permanent outflow approaches the continental shelf. This specific oceanographic phenomenon is an example of the stagnation in line flows studied in fluid mechanics. A brief summary of these flows will help to define the objective of this work and to place it in perspective with other branches of that discipline.
Following the classic boundary-layer theory for the solution of the two-dimensional incompressible steady flow of a homogeneous fluid near a stagnation point at a rigid boundary, it is convenient to divide the flow into the potential (irrotational) and the viscous (with nonzero vorticity) regions (Fig. 1). The flow vψ in the irrotational region is known to be described by the streamfunction ψ = kxy, where k is a positive constant and vψ = −k × gradhψ [see also other possible streamfunctions in, e.g., Peregrine (1981), Dorrepaal (1986), and Pedlosky (1987, p. 94)]. For the viscous region the governing Navier–Stokes equations can be reduced to an ordinary differential equation leading to similarity solutions. That solution was obtained numerically by Hiemenz (1911) and improved upon by Howarth (1935) [see, e.g., Batchelor (1970, chap. 5.5); Schlichting (1979) chap. V); Churchill (1988, chap. 14)]. Since that time, two-dimensional steady flow with a stagnation point has been referred to as “Hiemenz flow.” A similar division may be adopted in the oceanographic context. An oceanic jet approaching a continental slope has smaller turbulent eddy coefficients in the open ocean region than in the coastal region over the continental shelf. We will be primarily concerned with the changes experienced by such an oceanic free jet as it approaches the coast seaward of the 200-m isobath. Two important assumptions can therefore be justified for this situation. The first is that, as mentioned above, the flow is considered inviscid. The second is that the bottom depth is everywhere large enough to allow geostrophic velocities referred to a horizontal layer (e.g., at 200 m) of nearly zero motion to be computed from the density field. Unlike the theoretical jet studied by Hiemenz, oceanic jets cannot in general be considered irrotational. On the one hand, oceanic jets have large vertical shear and therefore a large horizontal component of relative vorticity. On the other hand, the vertical component of the relative vorticity even in straight impinging jets (i.e., ones with zero curvature vorticity) may be large due to the horizontal shear (shear vorticity).
The interest in quasi two-dimensional impinging flows having vorticity lies in the fact that the vorticity component parallel to the streamlines is susceptible to amplification by stretching and thus may be intensified near the boundary. The first studies of impinging two-dimensional flows with vorticity in the inviscid region were made in the context of the aeronautical sciences (Ferri and Libby 1954). Stuart (1959), after the studies of Li (1955, 1956) and Glauert (1957), considered the influence of uniform vorticity perpendicular to the plane of the two-dimensional flow of the fluid in the viscous region (such a fluid flow undergoes no stretching). Rott and Lenard (1959) and Kemp (1959) studied the same phenomenon but for axisymmetric stagnation flows. The effects of vorticity having a particular orientation susceptible to amplification by stretching when transported into the stagnation point boundary (turbulent) layer were described using a mathematical model by Sutera et al. (1963) and Sutera (1965). They demonstrated the possibility of vorticity amplification by stretching in quasi two-dimensional stagnation flow when the vorticity scale is larger than a certain neutral scale.
With the help of the above comments concerning general stagnation jets we are in a position to define this work as the study of the changes that a free jet with vorticity experiences when it approaches a solid boundary. Therefore, this work deals with the so-called external flow, that is, outside the viscous (turbulent) stagnation region, but such an external flow having vorticity. Furthermore, the geophysical context provides three new features that have not been addressed previously: namely, rotation, stratification, and a sloping boundary. To our knowledge the first work to address some consequences of the stagnation point region of a water jet in rotating fluid mechanics was that by Whitehead and Miller (1979) and later by Whitehead (1985a) in a laboratory simulation of the western Alboran gyre (WAG) in the Alboran Sea (Western Mediterranean) and the uplift near the Strait of Gibraltar. They pointed out that the growth of the gyre in their simulations may be due to the position of the stagnation point of the jet relative to coastal features at the point where the jet impinges on the African coast. They were unable to identify the existence of any theoretical or experimental studies of the properties of stagnation point flows in baroclinic, rotationally influenced jets. Whitehead considered the deflection by a vertical wall of a jet in a rotating fluid of constant density lying over a motionless fluid of slightly higher density. Using a momentum integral and considering a zero- as well as a constant-potential-vorticity jet, Whitehead was able to determine the relative size of the two branches of split flow as a function of the angle of incidence of the external jet.
The aim of this work is to characterize from observations the density and velocity changes experienced by a free, nearly geostrophic jet when it impinges on a sloping coastal boundary. We will therefore focus, as mentioned above, on the oceanic jet seaward of the continental shelf. We are aware that this topic is too wide and too complex to be resolved in one study. Consequently, our attempt is to describe the main hydrographic and kinematic properties that we have found to be clearly present in the impinging process. Therefore, the dynamical properties are excluded in this work. The specific case chosen in this study is the same one that motivated Whitehead’s studies, that is, the stagnation in line flow of the Atlantic current in the Alboran Sea. The description of the observations (section 2) is based on the density as well as on the vorticity and deformation fields. Due to the presence of a sloping coastal boundary, the changes in the jet’s structure occur in both the horizontal and vertical directions, that is, the changes are of a three-dimensional nature. Section 3 is devoted to the discussion and conclusions, and section 4 to the summary.
2. A case study: The Atlantic current in the Alboran Sea
The Atlantic current impinging on the African coast in the Alboran Sea is probably one of the best examples in nature of a quasi-steady jet impinging on a coast. Another possible example is the flow exiting the Tsugaru Strait north of Japan (Conlon 1982; Kawasaki and Sugimoto 1984). In both cases the geographic permanence of a jet is assured by a strait’s outflow. The knowledge of the physical oceanography of the Alboran Sea, although by no means complete, is one of the most extensive in the entire Mediterranean. The reader is referred to Parrilla and Kinder (1987) for a general description, Bucca and Kinder (1984) for meteorological effects, Gascard and Richez (1985) and Parrilla et al. (1986) for water masses, La Violette (1986) for short-term variability, Heburn and La Violette (1990) for gyre’s variability, Perkins et al. (1990) for the Atlantic inflow, Viúdez et al. (1996a, 1996b) for the gyre’s structure and ageostrophic motion, and Vázquez-Cuervo et al. (1996) for altimeter data analysis (and references therein).
The present data consists of 134 CTD stations (Δx ≈ 30 km, Δy ≈ 18 km) acquired in the Alboran Sea in September 1992 [described in detail in Viúdez et al. (1996a)]. Partial use is made of altimeter data [ERS1 (European Remote Sensing satellite) and TOPEX/Poseidon (ocean TOPography EXperiment)] in the discussion section in order to address the long-term temporal variability of the impinging process.
It has been the custom for ease of reference (e.g., La Violette 1986; Viúdez et al. 1996a) to use the term Atlantic jet (AJ) for the flow that, entering through the Strait of Gibraltar, usually impinges on the African coast at 3°40′W (Fig. 2). This flow then deflects to the left (facing downstream) and enters the eastern Alboran basin and then into the Algerian basin. The term WAG refers to the closed circulation including the flow that is deflected to the right at the stagnation point at the African coast. Strictly, the AJ and WAG are not two independent physical systems since there is water interchange between them (Lanoix 1974; Viúdez et al. 1996a). Since, in fact, they share a common density front, the current associated with this density front is here referred to as the impinging current, impinging jet, or simply the current, in order to distinguish it from the AJ (the impinging current being wider than the AJ). Using this terminology, the current impinges on the sloping African coast and is deflected and divided into two branches: the AJ, or left branch, and the WAG current, or right branch. Note that the horizontal and vertical scales of the African continental shelf in the impinging region (∼15 km and 200 m, respectively; see, e.g., Fig. 3) are of the same order as the horizontal and vertical scales of the Atlantic jet (or the impinging current). Thus, we expect that the sloping of the shelf influences the impingment process. Furthermore, since the width of the shelf is comparable to the meridional distance between CTD stations (∼18 km), we expect also that this influence may be detectable in the hydrographic data. Other possible features associated with the impinging process, but having spatial scales smaller than the spacing between CTD stations, could be missed or aliased with the present data resolution.
a. Hydrography
The hydrographic data can give us information about the splitting of the impinging current on the African coast (Fig. 3). Consider the σθ distribution on a cross-front vertical section of the current before its division into two branches (Fig. 4). Superimposed on this distribution we have indicated with dots the σθ values, between 50 and 150 m, found at the two CTD stations closest to the coast after the flow division (station E1 and G1 in Fig. 3). If, for this distance between CTD stations (∼30 km) at this depth (∼100 m), vertical displacement and diabatic terms in the conservation of potential density are neglected, the curves LW and LE linking these dots separate water that is to be deflected eastward (line LE) from the water to be deflected westward (line LW). The deflection of the fluid between line LW and LE cannot be determined from the CTD data. This amount of missing water, which produces the uncertainty in the determination of the direction of deflection, flows between the coast and the closest CTD stations offshore; that is, this flow cannot be resolved for the given location of the CTD stations. Note that due to the slope of the coast, this uncertainty decreases with depth. Line LWE, defined as equidistant to LW and LE in the horizontal, is therefore an approximation to the intersection of the stagnation streamsurface of the impinging current with the plane of vertical section 3. This line LWE represents, therefore, the division of the impinging current into the AJ and the WAG. Two important results can be drawn from this flow division. The first one is that the AJ flows eastward associated with a density front that is basically confined to the first 150 m (Fig. 5c), while the WAG water flows westward with the largest density gradients below 100 m (Fig. 5a). Note also the reversed baroclinicity [the slopes of isopycnals change sign with depth (Onken 1990)] in vertical section F (Fig. 5b) in between stations F4 and F2. The consequences that this reversed baroclinicity have in the vertical shear will be described in the next section.
The second result is that LWE has some tilt. This shows that the stagnation streamsurface does not lie in a vertical plane but tilts in the direction defined by the density gradients, the tilting tending to be diapycnal rather than isopycnal, that is, opposite to the vertical tilting of isopycnals. This tilting of the stagnation streamsurface is assured, in spite of the rather coarse alongcoast CTD spacing, by the fact that the CTD station F3 (fortunately) intersected the stagnation streamsurface itself. The approximate slope between 80 and 150 m is estimated to be α ≈ 70 m/30 km ≈ 5 × 10−3. As a consequence, offshore parcels in the upper part of a vertical column of water (denoted by PU in Fig. 4) intersecting LWE deflect eastward, while lower water parcels (PL) in the same vertical column of water deflect westward. As a consequence of the motion of PU and PL, the horizontal component of the vorticity normal to the vertical plane in Fig. 4 for a water parcel in between PU and PL (in the same water column) is positive (i.e., it points northward). Since the impinging current is southward, it is expected that water parcels on the stagnation streamline have negative streamwise vorticity (because the projection of the aforementioned vorticity along the velocity direction is negative).
The hydrographic data can also give us information about the modification of the density field in the AJ as it approaches the coast. Since to a first approximation the circulation can be considered horizontal, the existence of a sloping coast implies that, as a water column approaches the coast, the horizontal distance to the coast is larger for upper water parcels in the column than for lower water parcels. Therefore, changes in the horizontal distance to the coast for vertically separated distributions of potential density (σθ), as an approximate conservative property, is a way to infer possible coastal influences. Isolines of σθ at different horizontal planes are shown in Fig. 6. The values of σθ on each isoline couple correspond to those at four horizontally fixed locations (points A1 to A4) at a given depth in the AJ. These results show that the horizontal location of the isolines is approximately constant with depth from points A1 and A4 to ∼3°40′W, in such a way that the isolines fit in a narrow band. On the other hand, when the AJ curves southward toward the African coast and then turns eastward toward the eastern Alboran basin, the isolines at different depths no longer remain parallel. Potential density isolines in the same water column have, therefore, different orientations. We think this is a consequence of the different horizontal distance to the coast for surface and deep water elements in the AJ, that is, a consequence of the sloping coastal boundary. The kinematical manifestation of all the above features is presented in the next section.
b. Kinematics
In this section we show the main kinematic characteristics of the impinging region based on the vorticity and deformation fields. As part of the density data postprocessing, we obtained an adjusted, dynamically balanced three-dimensional velocity field (Viúdez et al. 1996b) through the dynamical assimilation of the density data into a PE model (Haney 1974, 1985) using the digital filter initialization technique (Lynch and Huang 1992). This method provided a dynamically balanced density and three-dimensional velocity field at the grid points of a 50 × 40 × 48 regular matrix (δx ≈ 9.3 km, δy ≈ 6.7 km, δz = 4 m) covering the entire Alboran Sea. This assimilation procedure, although not essential for most of the results presented here, provides a more accurate estimate of the velocity field than geostrophic.
The notation used in the remainder of the paper follows that in Gurtin (1981) and Truesdell (1991). We have adopted this notation instead of the Gibbsian dyadic notation (e.g., Morse and Feshbach 1953; Godske et al. 1957; Tai 1992), more common in hydrodynamics, because it is the one most commonly used in continuum mechanics, especially in the study of deformation. In particular a tensor
1) Curvatures and backing
The backing ΘZ is a measure of the change in direction of the velocity field along the vertical. The backing distribution (Fig. 7c) shows that ΘZ < 0 upstream far from the impinging region, but ΘZ > 0 upstream, and ΘZ < 0 past the stagnation point. Among these different areas characterized by different signs in ΘZ, probably the most relevant one is where ΘZ < 0, far upstream, because of the large velocities in the AJ. This importance can be easily visualized by looking at the streamlines at different depths in the AJ (Fig. 9). It is clearly observed how deep streamlines, “initially” aligned in the vertical with the surface ones in the AJ, are deflected to the left (facing downstream) “before” the surface ones.
In meteorology and oceanography the backing
2) Vorticity
The kinematic properties presented above are significant for several reasons. First, they represent a quantitative description of the three-dimensional velocity field in the impinging region. As such, they provide quantitative and meanful data for testing theories or models of the impinging process. One could, in principle, have done this description directly by simply mapping the velocity at different levels. Such a method, however, would be largely imprecise and only qualitative in nature. From the already mentioned non-Galilean invariance of these kinematic quantities, it is clear that they have no dynamical significance. Since the impinging jet is characterized by large changes of opposite sign in speed divergence and direction divergence (the changes in horizontal divergence can be neglected in comparison with the changes in speed divergence or direction divergence), the impinging process could be considered a very good example of a conversion between speed divergence and direction divergence and be studied using the interchange equations proposed by Hollmann (1958). However, such a treatment would also be considered a description and not an explanation of the changes in speed divergence and direction divergence (see equivalent conclusions regarding the shear and curvature vorticity equations in Viúdez and Haney 1996). The kinematic quantities presented above are also important because of their connection to certain experimental methods. For example, a satellite thermal image may be used to infer the streamline curvature field but, in general, not the velocity field. From the trajectory of a single isopycnal float we can measure the curvature vorticity, but not vorticity itself. We will see in the next two sections that many of the properties of the velocity field in this study, including the deformation, can be deduced from the single fact that the acceleration field is divergent in the impinging region.
3) Rotation and deformation
It should be noted that, as it is usual practice in oceanography, the velocity field used in the computation of WK is referenced to the rotating earth. Thus, WK is in fact a relative vorticity number. If velocity is instead referenced to a nonrotating system absolute vorticity should replace the relative vorticity in (11). The absolute vorticity number so obtained is one order of magnitude larger than the relative WK and is always positive in the study domain, with a relative extremum, in this case a maximum, in the impinging region. This shows the large effect that the solid (nondeformable) coastal boundaries have in communicating their own rotation to the oceanic flow.
4) Stretching
The study of stretching may be accomplished by means of the principal stretchings and principal axes of stretching. The reader is referred to Chadwick (1976), Gurtin (1981), and Truesdell (1991) for a detailed exposition of these topics. The principal stretchings are the proper numbers of the symmetric tensor
3. Discussion and conclusions
The results concerning the impinging process may be summarized by the sketch in Fig. 17. In the horizontal we distinguish between several areas labeled from I to IV. The first area (I) is upstream, far from the impinging region. In this area the horizontal flow is not normal to the coast, the stagnation streamlines have relatively large horizontal curvature, and the backing is negative (ΘZ < 0). When the flow approaches the coast and arrives at the impinging region proper, there is an upstream impinging region (area II) where the flow has speed convergence (δυh/δh < 0), and therefore diffluence (ΘN > 0) and speed deceleration (dυh/dt ≈ υhδυh/δsh < 0 in the steady state). This region is characterized by positive backing (ΘZ > 0) and, therefore, by negative streamwise pseudovorticity (−υhΘZ < 0). Past the impinging region, the two horizontal flow branches (areas III and IV) experience speed divergence (δυh/δsh > O), therefore confluence (ΘN < 0) and speed acceleration (dυh/dt ≈ υhδυh/δsh > 0 in the steady state). These two areas (III and IV) are characterized by negative backing (ΘZ < 0), and therefore by positive streamwise pseudovorticity (−υhΘZ > 0). While in area I the magnitude of the deformation and rotation are similar (WK ≈ 1), in areas II, III, and IV, which correspond to the upstream and past the stagnation point regions, the magnitude of deformation exceeds the magnitude of rotation (WK < 1). This fact is related to the divergence of the acceleration field. As the current approaches the coast there occurs a rotation (relative to the flow direction) of the stretching axes of the fluid element in such a way that they become almost parallel and normal to the coast in the impinging region.
The main general result of the present study is that, if it were not clear enough, there is an interaction between the current and the coast; that is, the current really impinges on the coast (for a rotating observer) and is deflected by the coast. Whereas, after looking at the satellite image (Fig. 2) and the density distribution (Fig. 3), our intuition tells us that this interaction is evident:the quantitative analysis of vorticity and deformation that we have done is the objective way of expressing this fact and describing how this coast–jet interaction takes place. This impinging process pattern in the Alboran Sea, while it appears to be the most frequent one, is not the only one possible. An interesting different case was observed by García-Lafuente et al. (1997), in which the AJ was “prematurely” deflected to the east by the presence of a small-scale cyclonic eddy located in the impinging region.
To first order the AJ impinging process is considered a quasi-steady process. That is, the temporal scale associated with the water elements passing through the impinging region (about one day) is considered to be small compared to the temporal scale of the impinging process itself. Probably the shorter periodic phenomenon that can substantially affect the impinging process is the quasiperiodic 9-day oscillation of the AJ reported by Perkins et al. (1990). Large-scale low-frequency variability in the Alboran Sea has been studied only from remote sensing data (Heburn and La Violette 1990; Vázquez-Cuervo et al. 1996). A time sequence of a part of the semiannual cycle of residual sea level in the Alboran Sea is shown in Fig. 18. The semiannual cycle was calculated by fitting at each grid point a semiannual harmonic to the merged TOPEX/Poseidon ERS1 altimeter data from October 1992 to December 1993. This dataset was processed by fitting the TOPEX/Poseidon orbit to the ERS1 orbit [which reduces substantially the orbit error of the ERS1 data; see Ayoub et al. (1997) for more information on the data processing]. It was found that in the Alboran Sea the annual and semiannual cycle each account for nearly 50% of the resolved variability. Clearly visible is an eastward propagation of the semiannual wave. It therefore appears that the impinging jet process could also have relevant annual and semiannual components. However, it is important to note that the total amplitude of the temporal variability of sea level (∼10 cm) is smaller than the mean sea level anomaly in the WAG–AJ system (∼25 cm, e.g., Perkins et al. 1990). Thus, for example, the meridional currents at the coast near 4°W are always directed toward the coast (southward) even though the disturbance currents inferred from the sea level gradients in Fig. 18 show a reversal.
In order to compare the impinging process in the Alboran Sea with the analytical results of Whitehead (1985b) it is required to assign an asymptotic angle of incidence to the impinging current. Because the impinging current has streamline curvature, there is no objective way for providing such an angle of incidence. Assuming that this angle is taken as zero, that is, the impinging jet is consider to be normal to the wall, Whitehead’s results predict that the percent volume of fluid deflected to the right is 65% and 75% for the zero-and constant-potential-vorticity jets, respectively. The geostrophic transport function computed from the same data we have analyzed (Viúdez et al. 1996a, Fig. 13) shows that the deflected geostrophic transport is ∼2 Sv (Sv ≡ 1 × 106 m3 s−1) to the right (WAG), and ∼1 Sv to the left (AJ). This corresponds to 66% of volume deflected to the right, a result that is consistent with Whitehead’s theory in spite of the different jet configurations and different wall geometry (vertical vs sloping coastal boundary).
On the other hand, Whitehead and Miller (1979) and Whitehead (1985a), based on their laboratory simulation of the WAG, pointed out that the growth of the gyre may be due to the position of the stagnation point of the jet relative to coastal features. In their simulations the growing of the gyre was stopped by the basin geometry. However, since the WAG is rarely found to occupy the entire upper layer of the western Alboran basin [we know of very few cases when this happened (Lanoix 1974; Cano 1978)], it appears that, for some reason, these laboratory experiments exaggerated the effect of the impinging process on the WAG circulation. It should also be noted that in the laboratory experiments with approximate coastal geometry (Whitehead and Miller 1979) it was required to smooth Cape Tres Forcas (the cape located on the African coast at 3°W) from the coastal geometry in order to develop a gyre.
The results concerning the impinging process, based on CTD data, show that, since the stagnation surface is tilted in the vertical, the magnitude of the streamwise vorticity increases due to the vertical shear in the across stream direction with the upper layer flowing to the left (facing downstream) and the lower layer flowing to the right. The impinging process described here is therefore a 3D feature. We, however, consider that no further conclusions concerning the remote origin of the water on both sides of the stagnation surface can be made from the CTD results presented here. Two main effects prevent us from this, namely, (i) the discrete CTD sampling and (ii) the diabatic and frictional effects that prevent surfaces of constant potential density and surfaces of constant potential vorticity from being material surfaces.
We would like to emphasize that the results presented here have been concerned with the so-called external flow and not with the turbulent flow in the viscous region. Although we have borrowed this flow division from the more general impinging process in fluid mechanics without proving it totally (because of the lack of measurements over the continental shelf), some results presented here (especially those concerning the irrotational characteristics of the impinging flow) point out that such a division may be acceptable in the oceanographic context. As a further consequence, since the impinging current has streamwise vorticity, it would be interesting to verify experimentally if vorticity amplification and an increase of turbulence occur over the African continental shelf, as it occurs in quasi two-dimensional impinging (nongeophysical) flows. Another possible phenomenon associated with an impinging process that may deserve further study in the oceanic case is the existence of hydraulic jumps, or standing waves, (e.g., Watson 1964; Higuera 1994; Fedorov and Melville 1996). If such a hydraulic jump exists, it could be observed along the African coast.
Impinging jet processes have received little attention in physical oceanography. For example, the study about the opposite phenomenon, that is, coastal current separation, is much more developed (e.g., Ou and De Ruijter 1986; Page 1987; Signell and Geyer 1991; Klinger 1994). We think that this is because the impingment process, although it may occur frequently in the ocean, is not usually a steady process (the temporal evolution of an eddy forced against a boundary has been studied by Shi and Nof 1993). Our main objective has been therefore to study, in a broad way, the main kinematic characteristics of an impinging jet in the Alboran Sea, and to relate this process with other disciplines of fluid mechanics. Since the presence the Atlantic jet is assured by the location of the Strait of Gibraltar, the impingment process in the Alboran Sea case is, concerning the steadiness, a privileged one.
4. Summary
We have described an oceanic impinging jet process in terms of its density, curvatures, backing, vorticity, and deformation characteristics. In the impingment region the acceleration field is divergent, which is related to the fact that the magnitude of deformation is larger than the magnitude of rotation. It also has been found that the stagnant streamsurface does not lie in a vertical plane but tilts in the opposite direction to the vertical tilting of isopycnals. The flow upstream of the stagnation point is characterized by backing, speed convergence, diffluence, and negative streamwise vorticity. The flow past the stagnation point is characterized by veering, speed divergence, confluence, and positive streamwise vorticity. The stretching axes of a fluid element approaching the coast experience a rotation relative to the flow direction in such a way that they become almost parallel and normal to the coast in the impinging region. Only in a narrow area of that impinging region can the current be considered irrotational.
Acknowledgments
The authors greatly appreciate a postdoctoral grant from the Ministerio Español de Educación y Ciencia and support from the Office of Naval Research (Code 322). N. Ayoub and P. Y. Le Traon are thanked for providing the altimetric dataset. One of the authors (J.V.-C.) is under contract with the National Aeronautics and Space Administration.
REFERENCES
Ayoub, N., P. Y. Le Traon, and P. De Mey, 1997: A description of the Mediterranean surface variable circulation from combined ERS-1 and TOPEX/POSEIDON altimetric data. J. Mar. Syst., in press.
Batchelor, G. K., 1970: An Introduction to Fluid Dynamics. Cambridge University Press, 615 pp.
Bjørgum, O., 1951: On Beltrami vector fields and flows. Part I. Universitet J. Bergen, Naturvitenskapalig rekke, No. 1, 1–85.
Brandes, E. A., R. P. Davies-Jones, and B. C. Johnson, 1988: Streamwise vorticity effects on supercell morphology and persistence. J. Atmos. Sci.,45, 947–963.
Bucca, P. J., and T. H. Kinder, 1984: An example of meteorological effects on the Alboran Sea gyre. J. Geophys. Res.,89, 751–757.
Cano, N., 1978: Hidrología del Mar de Alborán en primavera-verano. Bol. Inst. Esp. Oceanogr.,248, 51–66.
Chadwick, P., 1976: Continuum Mechanics. Wiley, 174 pp.
Churchill, S. W., 1988: Viscous Flows: The Practical Use of Theory. Butterworth, 602 pp.
Conlon, D. M., 1982: On the outflow mode of the Tsugaru Warm Current. La Mer,10, 60–64.
Davies-Jones, R., 1984: Streamwise vorticity: The origin of updraft rotation in supercell storms. J. Atmos. Sci.,41, 2991–3006.
——, 1991: The frontogenetical forcing of secondary circulations. Part I: The duality and generalization of the Q vector. J. Atmos. Sci.,48, 497–509.
Dhanak, M. R., and J. T. Stuart, 1995: Distorsion of the stagnation-point flow due to cross-stream vorticity in the external flow. Philos. Trans. Roy. Soc. London, A, 352, 443–452.
Dorrepaal, J. M., 1986: An exact solution of the Navier-Stokes equation which describes non-orthogonal stagnation-point flow in two dimensions. J. Fluid Mech.,163, 141–147.
Ericksen, J. L., 1960: Tensor fields. Handbuch der Physik III/I, S. Flügge, Ed., Springer-Verlag, 794–858.
Fedorov, A. V., and W. K. Melville, 1996: Hydraulic jumps at boundaries in rotating fluids. J. Fluid Mech.,324, 55–82.
Ferri, A., and P. A. Libby, 1954: Note on an interaction between the boundary layer and the inviscid flow. J. Aeronaut. Sci.,21, 130.
García-Lafuente, J., N. Cano, M. Vargas, J. P. Rubín, and A. Hernández-Guerra, 1997: Evolution of the Alboran Sea hydrographic structures during July 1993. Deep-Sea Res., in press.
Gascard, J. C., and C. Richez, 1985: Water masses and circulation in the western Alboran Sea and in the Straits of Gibraltar. Progress in Oceanography, Vol. 15, Pergamon Press, 157–216.
Gluert, M. B., 1957: The boundary layer in simple shear flow past a flat plate. J. Aeronaut. Sci.,24, 848–849.
Godske, C. L., T. Bergeron, J. Bjerknes, and R. C. Bundgaard, 1957:Dynamic Meteorology and Weather Forecasting. Amer. Meteor. Soc. and Carnegie Institute, 800 pp.
Gola̧b, S., 1974: Tensor Calculus. Elsevier, 371 pp.
Gurtin, M., 1981: An Introduction to Continuum Mechanics. Academic Press, 265 pp.
Haney, R. L., 1974: A numerical study of the response of an idealized ocean to large-scale surface heat and momentum flux. J. Phys. Oceanogr.,4, 145–167.
——, 1985: Midlatitude sea surface temperature anomalies: A numerical hindcast. J. Phys. Oceanogr.,15, 787–799.
Hawthorne, W. R., 1951: Secondary circulation in fluid flow. Proc. Roy. Soc. London Ser. A, 206, 374–387.
——, and M. E. Martin, 1955: The effect of density gradient and shear on the flow over a hemisphere. Proc. Roy. Soc. London Ser. A, 232, 184–195.
Heburn, G. W., and P. E. La Violette, 1990: Variations in the structure of the anticyclonic gyres found in the Alboran Sea. J. Geophys. Res.,95, 1599–1613.
Hiemenz, K., 1911: Die Grenzschicht an einem in den gleichförmigen Flüssigkeitsstrom eingetauchten geraden Kreiszylinder. Ph.D. dissertation, Göttingen and Dingl. Polytech. J.
Higuera, F. J., 1994: The hydraulic jump in a viscous laminar flow. J. Fluid Mech.,274, 69–92.
Hollmann, G., 1958: Die Krümmungs- und die Scherungs-Vorticitygleichung; Die Richtungs- und die Geschwindigkeits-Divergenzgleichung. Beitr. Phys. Atmos.,30, 254–267.
Hoskins, B. J., 1975: The geostrophic momentum approximation and the semi-geostrophic equations. J. Atmos. Sci.,32, 233–242.
Howarth, L., 1935: On the calculation of steady flow in the boundary layer near the surface of a cylinder in a stream. Aero. Res. Counc., Lond., R&M no. 1632.
Kawasaki, Y., and T. Sugimoto, 1984: Experimental studies on the formation and generation processes of the Tsugaru Warm Gyre. Ocean Hydrodynamics of the Japan and East China Seas, T. Ichiye, Ed., Oceanogr. Series, No. 39, Elsevier, 225–238.
Kemp, N. H., 1959: Vorticity interaction at an axisymmetric stagnation point in a viscous incompressible fluid. J. Aerosp. Sci.,26, 543–544.
Klinger, B. A., 1994: Inviscid current separation from rounded capes. J. Phys. Oceanogr.,24, 1805–1811.
La Violette, P. E., 1986: Short term measurements of surface currents associated with the Alboran Sea during ¿Donde Va? J. Phys. Oceanogr.,16, 262–279.
Lakshminarayana, B., and J. H. Horlock, 1973: Generalized equations for secondary vorticity using intrinsic co-ordinates. J. Fluid Mech.,59, 97–115; Corrigendum, 226, 661–663.
Lanoix, F., 1974: Projet Alboran, Etude hydrologique et dynamique de la Mer d’Alboran. Tech. Rep. 66, NATO, Brussels, 39 pp. plus 32 figs. [Available from NATO Subcommittee on Oceanographic Research, Copenhagen, Denmark.].
Li, T. Y., 1955: Simple shear flow past a flat plate in an incompressible fluid of small viscosity. J. Aeronaut. Sci.,22, 651–652.
——, 1956: Effects of free-stream vorticity on the behaviour of a viscous boundary layer. J. Aeronaut. Sci.,23, 1128–1129.
Lynch, P., and X.-Y. Huang, 1992: Initialization of the HIRLAM model using a digital filter. Mon. Wea. Rev.,120, 1019–1034.
Marris, A. W., 1964: Generation of secondary vorticity in a stratified fluid. J. Fluid Mech.,20, 117–181.
——, and S. L. Passman, 1968: Vector fields and flows on developable surfaces. Arch. Ration. Mech. Anal.,31, 29–86.
Masotti, A., 1927: Decomposizione intrinseca del vortice e sue applicazioni. Inst. Lombardo Sci. Lett. Rendiconti,60(2), 869–874.
Morse, P. M., and H. Feshbach, 1953: Methods of Theoretical Physics. Part I. McGraw-Hill, 997 pp.
Onken, R., 1990: The creation of reversed baroclinicity and subsurface jets in oceanic eddies. J. Phys. Oceanogr.,20, 786–791.
Ou, H. W., and W. P. M. De Ruijter, 1986: Separation of an inertial boundary current from a curved coastline. J. Phys. Oceanogr.,16, 280–289.
Page, M. A., 1987: Separation and free-streamlines flows in a rotating fluid at low Rossby number. J. Fluid Mech.,179, 155–177.
Parrilla, G., and T. H. Kinder, 1987: The physical oceanography of the Alboran Sea. Rep. Meteor. Oceanogr.,1, 143–184.
——, ——, and R. H. Preller, 1986: Deep and intermediate Mediterranean water in the western Alboran Sea. Deep-Sea Res.,33, 55–88.
Pedlosky, J., 1987: Geophysical Fluid Dynamics. 2d ed. Springer-Verlag, 710 pp.
Peregrine, D. H., 1981: The fascination of fluid mechanics. J. Fluid Mech.,106, 59–80.
Perkins, H., T. Kinder, and P. E. La Violette, 1990: The Atlantic inflow in the western Alboran Sea. J. Phys. Oceanogr.,20, 242–263.
Rott, N., and M. Lenard, 1959: Vorticity effect on the stagnation-point flow of a viscous incompressible fluid. J. Aerosp. Sci.,26, 542–543.
Sadeh, W. Z., and H. J. Brauer, 1980: A visual investigation of turbulence in stagnation flow about a circular cylinder. J. Fluid Mech.,99, 53–64.
Schlichting, H., 1979: Boundary-Layer Theory. McGraw-Hill, 817 pp.
Schouten, J. A., 1954: Ricci-Calculus. An Introduction to Tensor Analysis and Its Geometrical Applications. 2d ed. Springer-Verlag, 516 pp.
Scorer, R. S., 1978: Environmental Aerodynamics. Ellis Horwood, 488 pp.
——, and S. D. R. Wilson, 1963: Secondary instability in steady gravity waves. Quart. J. Roy. Meteor. Soc.,89, 532–539.
Shi, C., and D. Nof, 1993: The splitting of eddies along boundaries. J. Mar. Res.,51, 771–795.
Signell, R. P., and W. R. Geyer, 1991: Transient eddy formation around headlands. J. Geophys. Res.,96, 2561–2575.
Speziale, C. G., 1987: On helicity fluctuations in turbulence. Quart. Appl. Math.,45, 123–129.
Squire, H. B., and G. Winter, 1951: The secondary flow in a cascade of airfoils in a nonuniform stream. J. Aeronaut. Sci.,18, 271–277.
Stuart, J. T., 1959: The viscous flow near a stagnation point when the external flow has uniform vorticity. J. Aerosp. Sci.,26, 124–125.
Sutera, S. P., 1965: Vorticity amplification in stagnation-point flow and its effect on heat transfer. J. Fluid Mech.,21, 513–534.
——, P. F. Maeder, and J. Kestin, 1963: On the sensitivity of heat transfer in the stagnation-point boundary layer to free-stream vorticity. J. Fluid Mech.,16, 497–520; Corregendum, 17, 480.
Tai, C.-T., 1992: Generalized Vector and Dyadic Analysis. IEEE Press, 134 pp.
Truesdell, C., 1953: Two measures of vorticity. J. Ration. Mech. Anal.,2, 173–217.
——, 1954: Kinematics of Vorticity. Indiana University Press, 232 pp.
——, 1991: A First Course in Rational Continuum Mechanics. Vol. 1. 2d ed. Academic Press, 391 pp.
——, and R. Toupin, 1960: The classical field theories. Handbuch der Physik III/I, S. Flügge, Ed., Springer-Verlag, 902 pp.
Vázquez-Cuervo, J., J. Font, and J. J. Martínez-Benjamín, 1996: Observations on the circulation in the Alboran Sea using ERS1 altimetry and sea surface temperature data. J. Phys. Oceanogr.,26, 1426–1447.
Viúdez, A., and R. L. Haney, 1996: On the shear and curvature vorticity equations. J. Atmos. Sci.,53, 3384–3394.
——, J. Tintoré, and R. L. Haney, 1996a: Circulation in the Alboran Sea as determined by quasi-synoptic hydrographic observations. Part I: Three-dimensional structure of the two anticyclonic gyres. J. Phys. Oceanogr.,26, 684–705.
——, R. L. Haney, and J. Tintoré, 1996b: Circulation in the Alboran Sea as determined by quasi-synoptic hydrographic observations. Part II: Mesoscale ageostrophic motion diagnosed through density dynamical assimilation. J. Phys. Oceanogr.,26, 706–724.
Watson, E. J., 1964: The radial spread of a liquid jet over a horizontal plane. J. Fluid Mech.,20, 481–499.
Whitehead, J. A., Jr., 1985a: A laboratory study of gyres and uplift near the Strait of Gibraltar. J. Geophys. Res.,90, 7045–7060.
——, 1985b: The deflection of a baroclinic jet by a wall in a rotating fluid. J. Fluid Mech.,157, 79–93.
——, and A. R. Miller, 1979: Laboratory simulations of the gyre in the Alboran Sea. J. Geophys. Res.,84, 3733–3742.