1. Introduction
“Chaotic tides” sounds like an oxymoron. As the etymology of the word “tide,” being time, season, or period according to Webster's (1913) Revised Unabridged Dictionary implies, it has, from ancient times on (Pugh 1987), often been held to be synonymous with periodicity. Indeed, this regularity made tidal heights one of the earliest known and best predictable phenomena in physical oceanography, with relative accuracies of prediction often exceeding 90%. The predictive capacities of the corresponding tidal currents is, admittedly, less impressive, but is still often in excess of 50%. Less known, however, is the fact that spurious, but persistent, reports on the irregularity of the tides exist for over a century (Honda et al. 1908; Krümmel 1911). It is this irregularity in the tide that we here aim to identify as being possibly due to the chaotic response of certain coastal basins to the tide at the open sea.
Tidal predictions have traditionally been made for individual ports on the basis of previously observed tidal elevation records at those locations. Amplitudes and phases of a number of precisely known “tidal frequencies,” stemming from the gravitational potential determined by celestial mechanics, are estimated, which can then be used to sum the corresponding harmonic series and thus to predict the future occurrence of tides at that location; see, for example, the review by Godin (1991).
Although the gravitational (or tidal) potential carries only about 5 relevant, independent frequency components, which, through an expansion of the tidal potential in harmonic terms leads to perhaps 11 principal tidal components (Platzman 1971), in the shallow, coastal areas, local nonlinear effects lead to leakage of energy to higher harmonics and sum and difference (combination) frequencies. In this way tidal energy fills in entire spectral bands surrounding the principal components (Pugh 1987, p. 188). As there are, in principle, no difficulties in also resolving the amplitudes and phases of these combination frequencies, running harmonic analysis with 150 frequency components or more has become routine procedure. This empirical local analysis, however, does not automatically guarantee that the amplitudes and phases of these independently determined frequencies also show a spatially coherent and stationary picture. Quite the contrary.
A spatially coherent description of the tides can be obtained by solving the governing dynamical equations, the Laplace tidal equations (see, e.g., Cartwright 1977). In a realistic ocean domain this is achieved by numerical methods (Schwiderski 1980). Outside continental margins, tidal predictions for these principal components, particularly when constrained by observations from deep-sea tidal stations, are in fair agreement with observations obtained at different seaports and deep sea tidal stations (global root-mean-square difference between modeled and observed elevations of about 10 cm). With the advent of satellite altimetry providing tidal observations, this agreement has in recent years actually been dramatically improved (global root-mean-square difference of about 3 cm); see Andersen et al. (1995).
Except for these directly driven, principal components, together perhaps with their first few harmonics (Davies 1986; Lynch and Werner 1991), spatial coherency is cumbersome and agreement with locally observed tidal components poor, particularly in the shallow, coastal areas. Waves at combination frequencies are generated locally, get amplified, or damped, to a degree depending on the particular resonance properties of that locality, and do not persist as free waves so that they decorrelate quickly. Treating the nonlinearly generated compound waves locally as linearly independent (as harmonic, or Fourier analysis wants it) is therefore somewhat contorted but still defendable when the local nonlinear transfer of energy is frozen so that harmonics appear as the result of a stationary process. However, when analyzing different, independent datasets from the same location, variations in tidal amplitudes and phases often still occur, to the extent that some of the minor components appear, in fact, unresolvable (Gutiérrez et al. 1981; Godin 1991). Although such variations in locally estimated tidal amplitudes and phases are usually attributed to nonstationary effects due to wind (van Ette and Schoemaker 1966; Gutiérrez et al. 1981), this need not necessarily be its only source. Nonstationarity may also be due to nonlinearity of the hydrodynamic system itself (Pugh 1987), which may not only change the tidal elevation profile, by giving rise to superharmonics (overtides, which stay fixed in time), but may also modulate its amplitude, by giving rise to subharmonics. Finally, strong nonlinearity may provoke a cascade of such subharmonic bifurcations, giving rise to chaotic behavior.
Is there any observational basis for chaotic behavior of the tides, other than the aforementioned unresolvability, or noisiness, of tidal “constants”? In an analysis of the tides in Venice Lagoon, at the head of the Adriatic Sea, where the tides seem to pick up because of near-resonancy of the basin, Vittori (1992) observed that consecutive tidal maxima are highly irregular. She argued this to be indicative of low-dimensional chaos. Whether the low-order dynamics to which this is due is either inherited from the dynamics of the local wind fields or of a genuinely oceanographic nature is not clear. Similar changes in consecutive maxima also appear in long-term, tidal elevation records in the Wadden Sea basins (J. T. F. Zimmerman 2000, personal communication), which are again close to resonance (see Maas 1997, hereafter referred to as M). Frison et al. (1999) find evidence of nonlinear tides in U.S. coastal estuaries from the nontrivial shape of the attractor obtained in a diagram where observed elevation at some moment is plotted against that observed at some previous instant.
Surprisingly, direct observations of irregular oscillations, accompanying the tides, have been reported for over a century (see Krümmel 1911, 157–185; Defant 1961, p. 187). The suggestion that these “secondary oscillations” were actually related to the “primary,” tidal oscillation was firmly put forward in an impressive study by Honda et al. (1908), who had intensively studied the seas around Japan. This study eventually led to a complete classification of bays in terms of “periodic,” “quasiperiodic,” etc. (Honda et al. 1908; Nakano, 1932). In the absence of a dynamical framework, such as recently offered by the field of nonlinear dynamics, this classification was, however, not further interpreted, and these observations seem, if not completely forgotten, neglected in contemporary tidal literature. One reason for this might be that these secondary oscillations generally seem inconspicuous, as they consist of high-frequency, low-amplitude waves superimposed on a low-frequency, high-amplitude (primary) tide. Typical periods range from some minutes to several hours, and amplitudes do not exceed a few percent of the tidal range. However, an interesting twist is given to this interpretation by recent observations of such irregular tidal elevations in a Norwegian fjord (Golmen et al. 1994). By making simultaneous observations of the currents in the strait connecting the fjord to the sea they found that these irregular small-amplitude elevations were accompanied by irregular O(1) variations of the tidal current (see Fig. 1). As velocities are determined by elevation differences over the strait, by inference the velocity field can be amplified when, at any moment, the difference in tidal elevation also amounts to just a few percent of the elevation itself, thereby becoming of comparable magnitude to the difference between the (small amplitude) high-frequency elevation that is resonantly excited within the basin but is practically absent at sea. Such O(1) velocity variations are not only important for nautical reasons, but are clearly equally relevant for the flushing of the fjord: the exchange of water, sediments, and dissolved gases or nutrients with the connected sea. It is believed (LeBlond 1991) that this property will transcend the global relevance that both dissipation as well as resonance of tides in the coastal environment may have in setting the boundary conditions for the global tide (Garrett 1975), although Munk and Wunsch (1998) recently made the interesting suggestion that tides might be playing a key role in ocean circulation.
In order to illustrate these issues we will consider the resonant response of a short, deep basin having a sloping bottom, to the tide in the adjacent open sea to which it is connected by a narrow strait—a nonlinear Helmholtz resonator (see M). This choice of geometry differs from the mostly wide bays with their quarter-wave resonancies, encountered by Honda et al. (1908), but it ensures that the tide within the basin takes the simplest form possible, as it is governed by the pumping, or Helmholtz mode. This mode is the lowest and generally also the most important frequency in the basin's spectrum, and is characterized by a spatially uniform response. It can thus be represented by a single state variable (the excess volume of water, related to the free surface elevation), whose evolution is therefore described simply by an ordinary differential equation. Although nonlinearity can also be present in the frictional damping term (Zimmerman 1992), it is the nonlinearity in the restoring term that gives rise to multiple equilibria and chaos here. The latter nonlinearity is due to the sloping bottom (Green 1992) and can be understood as follows. The current through the strait is driven simply by the elevation difference over the strait. However, the elevation change within the basin, affected by the transport through the strait, clearly depends on the surface area that the incoming water has to cover, which is nonuniform with depth when the walls of the basin are not vertical. The time needed to produce a particular elevation change therefore depends on the preexisting sea level; see Fig. 2. In M its free, forced, and forced-and-damped response was discussed in the case where the tidal forcing is “pure,” that is, contains one frequency component only. In inviscid circumstances, a single frequency forcing can provoke a chaotic response in a small parameter range. However, when damping is added, the chaos fails to persist. Since the inclusion of damping terms is mandatory in any realistic physical context, this “Hamiltonian chaos” appears to have no physical relevance. It is one of the goals of the present paper to determine under what conditions tides, within basins of the kind considered here, can become chaotic, even in the presence of damping. Multiple frequency forcing seems to be a prerequisite for this. Therefore, in light of the observations discussed above, we will consider forcing with either two nearby frequencies, as is the case for a combination of lunar and solar tides, a “mixed” type of tide (see Defant 1961), or with a single low-frequency tide, accompanied by a small-amplitude, high-frequency, resonant perturbation, as is the case in a fjord.
In section 2 the nonlinear tidal Helmholtz resonator will be introduced. It represents the response of a bay to tidal variations at the open sea, to which it is connected by a narrow channel. Its governing ordinary differential equation will be derived. This section also introduces the Poincaré phase plane, which gives a more comprehensive way to follow its evolution than that obtained by direct numerical integration. The tidal response of the bay typically shows a gradual modulation of amplitude and phase, which can be captured by averaging the original governing equation (section 3). In section 4 this modulation equation is applied to the case that forcing is at two nearby frequencies. When these are close to resonance, the response of the basin may, under certain conditions, be chaotic. The case that a single, low-frequency tide is accompanied by a small-amplitude, high-frequency resonant perturbation is very similar to this, and is discussed in section 5. In section 6 the results will be summarized.
2. Nonlinear Helmholtz resonator
a. Poincaré plane, free response, and periodic forcing
To summarize, in this Poincaré plane, a strictly periodic oscillation will be represented by a single (fixed) point, a quasiperiodic oscillation by multiple points lying on a closed curve (limit cycle), and an aperiodic state (encountered in later sections) by an irregularly located set of points (strange attractor).
b. Qualitative effect of quasiperiodic forcing
The multiple equilibria near the Helmholtz resonance correspond to an amplified and a choked tidal response (see the two centers in Fig. 4d, at large and small volumes |x|, respectively). These states are stable and attracting when damping is present (Fig. 4d). Without damping, however, these states would retain their original distances to the fixed points (as the curves in Fig. 4b), and each state can be characterized by its energy level and period (with which it is traversed). The qualitatively different regimes, recognizable in Fig. 4b, are separated from each other by particular special orbits, the separatrices. One outer separatrix separates the outermost orbit from the other two, and one inner separatrix separates the banana-shaped orbit and small circular orbit. The period needed to traverse these separatrices approaches infinity, as the intersection of the two separatrices acts like a saddle. A state can creep up to this relative maximum in potential energy by exchanging some of its “kinetic” energy, allowing it to move forward along its orbit, in favor of “potential” energy.
When friction is introduced, it pulls the state down into one of the equilibria, except for the odd point sitting right at the saddle and the two orbits that limit exactly on the saddle point in forward time. Note that these orbits merge with the aforementioned separatrices as the friction decreases to 0. Apart from these exceptions, all initial states thus belong to one of two domains of attraction (amplification or choking). When one starts to weakly and slowly “shake” the system (i.e., the saddle and the separatrices) by introducing a large-period perturbation, a new &ldquo=uilibrium” may arise. This happens because it is clearly quite a sensitive matter near the separatrices to which equilibrium their state will recede. Therefore, when, due to shaking, the separatrices move a little, while the state is slowly receding to one equilibrium, it may find itself a little later in the attraction domain of the other. Making again its way to the other equilibrium, something similar may happen, and, in effect, the neighborhoods of the separatrices may trap the state.
This new “steady” state might turn out, however, to be chaotic. This is because, close to the separatrices, the period of the inviscid orbits (the modulation period of the Helmholtz oscillator) varies greatly, and thus, when the state of the system varies due to the weak forcing, neighboring states fastly diverge from each other.
The way to verify that such a new dynamical equilibrium state turns into existence under the addition of perturbative damping and extra forcing is by performing an energy budget study. For this, one picks one of the original inviscid orbits, characterized by a particular period, and one then evaluates, along that orbit, the net increase or decrease of energy over one period due to both the work done by the perturbative forcing as well as the energy dissipation by the perturbative damping term. Only when these two are in balance (and there is thus no net energy increase or decrease) will the point on average stay near the original, unperturbed orbit. The single orbits considered in the present study are the previously introduced separatrices (homoclinic orbits), along which the period turns to infinity. The function determining the net energy change on the homoclinic orbits is called the Melnikov function. This function is usually given the alternative interpretation of measuring the distance of the perturbed orbits leaving and approaching the saddle. A (traditional) search for its zeros—implying intersections of the two perturbed orbits so that points both belong to the sets of points approaching and leaving the (perturbed) saddle, and are therefore “stuck”—is thus equivalent to the requirement that there is no net change in energy content along its trajectory. Mathematically, all we can show for the moment is that certain points (that belong to this invariant set) get trapped near the unperturbed separatrices, and that they traverse that region in an unpredictable, chaotic way (section 4a). However, we lack the ability to rigorously show that this set is actually attracting (Wiggins 1990): Suggestions that it is follow from numerical analysis (section 4b).
Similar such phenomena appear at other resonances, like near ω ≈ 1/2 + ε, and in particular near ω ≈ 2 ± ε (see M).
3. Modulation equations
For the single-frequency forcing one can assume, without loss of generality, θ = 0 (which amounts to a rotation of our coordinate system). Its explicit presence is retained here however for later use, when, in the double-frequency forcing case, both F and θ will be varying on the slow timescale. Note that these modulation equations are the same as those found for the Duffing equation [in polar coordinates by the method of multiple scales, see Nayfeh and Mook (1979), and in both polar and Cartesian coordinates by the averaging method, see Guckenheimer and Holmes (1983)]. The extension to two forcing frequencies, discussed later, should thus have relevance to the quasiperiodically forced Duffing equation (Wiggins 1990; Yagasaki 1990, 1993), and extends it to the “double resonant” case of two nearby frequencies, both approximately equal to the natural frequency.
a. Comparison of exact and modulated systems
It is useful to understand the system of modulation equations because this local approximation has a strong correspondence with the Poincaré plot of the exact equation. To appreciate this, compare some of its numerically obtained solution curves with Poincaré plots of the exact equation (Fig. 6). Not only is the topology of fixed points and separatrices preserved but, indeed, so is much of their location (when giving proper notice to the rescaling factor ε, which is present in between these two systems). The modulation equations, however, have no restrictions on the region attainable in X–Y phase space, while the exact equations do. It is probably this restricted access of the original phase plane that makes a (formally) local solution perform so well in approximating its “global” dynamics. We will find that all of the behavior found near resonance in the original phase space is mimicked in the phase plane of the modulation equations (including, as we will see, the occurrence of chaos), which therefore forms a useful substitute.
b. Hamiltonian description
c. Homoclinic orbits
The motivation for studying the inviscid case (C = 0) is related to the fact that the modulation equations can then be solved explicitly. This offers the means to calculate Melnikov functions explicitly, once, at a later stage, weak (slow timescale) forcing and damping are added to the modulation equations. Zeros of the Melnikov functions determine the presence of invariant chaotic sets and suggest the presence of chaotic solutions. The conditions under which zeros of these functions appear will map out regions in parameter space where one may find “chaotic tides.”
4. Forcing at two nearly resonant frequencies
Thus, taking into account an additional solar, and therefore almost resonant, tidal forcing term gives a mechanism for the construction of chaotic solutions (see also Yagasaki 1990, 1993).
The derivation of the modulation equations is now identical to the derivation given for the single-frequency case in section 3 (now one averages over 2π/ω1 instead of 2π/ω) and the modulation equations are again of the form (9)–(10), or, in Cartesian coordinates, (11)–(12), but now with slowly modulating forcing F(T) = f(T)/ε3 and phase θ(T), Eqs. (27) and (28) replacing their constant counterparts (see Nayfeh and Mook 1979, section 4.4). The right-hand sides of the modulation equations (11)–(12) now depend on the slow time explicitly, and, in principle, this offers the opportunity for the occurrence of aperiodic solutions.
When the two components are of equal magnitude, f2 = f1, the phase is simply θ = (ΔΩ T + θ1 + θ2)/2, which redefines the effective frequency to be equal to the average frequency. The amplitude gradually changes sign over a long timescale f = 2f1 cos[(ΔΩ T + Δθ)/2]. The block version of this, f = 4f1(1/2 − Θ{cos[(ΔΩ T + Δθ)/2]}, is piecewise integrable. Here Θ(x) = 0 (1) for x < 0 (>0) denotes the Heaviside function. It has a saddle point on the X axis, whose position alternates with respect to the center of the X–Y phase plane with subsequent phases of the forcing. It is thus reminiscent of Aref's (1984) blinking vortex with its ensuing chaos. This suggests similar results for a sinusoidal forcing, which is confirmed by numerical analysis (see section 4b), although this cannot be substantiated analytically.
When the second component is but a small perturbation with respect to the first one f2/f1 ≡ δ ≪ 1, then f ≈ f1[1 + δ cos(ΔΩ T + Δθ)] and θ ≈ θ1 + δ sin(ΔΩ T + Δθ). Note that this occurs quite naturally in the context of M2 and S2 tides. In contrast to the previous case, this limit is more amendable to further analysis and will be addressed now. Without loss of generality we set θ1 = 0, fixing the origin of the “fast” time t, and Δθ = −π/2, fixing the origin of the “slow” time T, whence the amplitude f = f1[1 + δ sin(ΔΩ T)] and phase θ = −δ cos(ΔΩ T), so that, with f = ε3F, for small values of δ, apart from a factor 2, the forcing terms in (11) and (12) are approximated as F sinθ ≈ −δF cos(ΔΩ T) and F cosθ ≈ F[1 + δ sin(ΔΩ T)], respectively.
a. Perturbative second forcing component and Melnikov analysis
Neglecting for the moment the additional forcing term [last term in Eq. (30)], the effect of the damping is to draw trajectories inward across separatrices. Multiplying (30) with ∇H, the rate of change of the Hamiltonian is then obtained, on the separatrix (where K = −3
A (nondegenerate) zero of, for example, Mout(T0) corresponds to a transversal intersection of the “outer” stable and unstable manifolds of
In this paper we only consider the “classical way” to construct chaotic solutions, namely due to intersections of either the inner stable and unstable homoclinic orbits, or the outer ones. Other possibilities, like the intersection of the outer unstable manifold of
Note that both Cout and Cin vanish when the two basic frequencies are equal (i.e., ΔΩ = 0). In this case both Mout(T0) and Min(T0) are equal to M1 (for definition of M1,2,3 see appendix B), which does not depend on T0, and thus, in this case there can be no (transversal) intersections of the stable and unstable manifolds for the Poincaré map
b. Numerical solutions of modulation and exact equations
Proving that the modulation equations possess a chaotic invariant set in the case when the Melnikov function vanishes is as far as we go analytically. As for the (closely related) Duffing equation, more work is needed to show that this set becomes attractive in some parameter range (Wiggins 1990, 612–613). In the following we will only give some numerical support that the (periodically forced) modulation equations possess such a chaotic (strange) attractor. This, in turn, suggests that the original, quasiperiodically forced Helmholtz oscillator should likewise have a strange attractor in the corresponding parameter regime; a result that we confirm by numerical integration below.
Numerical integration of the modulation equations, (11)–(12), with a fourth-order Runge–Kutta scheme, employing a double-frequency forcing (23), shows that for two nearby frequencies within the frequency range over which multiple equilibria exist, the solution curves appear to be chaotic (Fig. 9). This is suggested both in the (slow) time domain T, by the strait's current speed Y (Fig. 9a), as well as in the phase space of current velocity Y versus excess volume X (Fig. 9b). (Recall that the original x and averaged quantities X are related by x ∝ εX, while the fast t and slow T timescales are related by t = ε−2T.) Here we take forcing amplitudes of the two external tidal components to be equal. One of these components is supposed to be at the (linear) resonance frequency, 1 (the Helmholtz frequency), while the second is at a slightly higher frequency (1.01). The latter frequency is setting the small parameter ε = 0.1. Within the resonance band the theory requires the forcing amplitudes to be small, O(ε3), so that we take them f1 = f2 = 10−3. This leads, from (27), to F(t) = f/ε3 = 2 cos[(ΔΩ T + Δθ)/2] in the modulation equations (12). In striking contrast to the monofrequency case, where the Hamiltonian chaos disappears with the introduction of even the smallest amount of damping (see M), the present calculations show that chaos survives the addition of damping. Commensurate with the theoretical requirements we need to take c ≤ O(ε2) and, in fact, take c = ε3 here. This amounts to δC = 0.1. The erratic solution curves suggest that the modulation equations may possess a strange attractor, which is indeed revealed by making a Poincaré plot, by sampling at the modulation period (Fig. 10). Its shape is very similar to the “Japanese attractor” (see Ruelle 1980), which has been encountered for the forced and damped Duffing equation. Notice, in particular, the familiar Cantor-like division of the attractor's branches.
Chaos in the modulation equations suggests the occurrence of chaos in the original equations (5) under double-frequency forcing (23). In the following numerical experiments, the same values for forcing amplitudes and frequencies, as well as damping, were taken as were used in the previous numerical integration of the modulation equations. In order to compare the slow modulation of the numerically computed orbits directly to those obtained from the modulation equations, we sample the former at the tidal period. So, in Fig. 11 Poincaré plots of damped (c = 0.001) flow rate y versus time t (panel a) and versus excess volume x (panel b) are shown. Sampling is done once every average period 2π/
5. Forcing at two widely separate frequencies
Consider first the case in which the perturbation is absent altogether, b = 0. Then, when damping is relatively weak because the frequency of the tide ε2 ≪ 1, the elevation within the basin will be able to follow the external tide: ζ ≈ Z. Therefore, in the presence of a high-frequency perturbation (0 ≠ b ≪ a), on the fast timescale the dominant component of the tide, Z, acts as a slow, quasi-adiabatic change of the mean depth of the basin. However, the basin depth (among other parameters) determines the Helmholtz frequency; see (1). Therefore, the frequency of the perturbation, assumed to be fixed in absolute measure, is slowly varying when scaled with this Helmholtz frequency. This will provoke the quasi-stationary response to drift along the response curve (see Fig. 5). When this frequency drift, induced by the apparent modulation of water depth, is large enough—covering the frequency range over which multiple equilibria exist, this will lead to a sequence of “catastrophes”: rapid changes in the range of the (near) resonant, high-frequency response, which is subject to periodical collapse and expansion. There will thus be a hysteretic change in the amplitude of the basin elevation. This change may, however, be chaotic because the moments in time at which the amplitude jumps can become randomized. This is perhaps so because the high-frequency response, if not rapidly constrained to a particular phase by strong damping, will still be in a transient adapting phase, recovering from the previous catastrophe, when it undergoes the next one.
It is ironic that this explanation, based on the possibility of undergoing a hysteretic curve requires fairly large, but not too large, damping magnitudes: if damping gets too small, the frequency range over which multiple equilibria exist becomes broader than the range of the modulating, detuned frequency, and the response is “stuck” to a particular branch; if, on the other hand, damping is too strong, the response curve has no multiple equilibria at all, again lacking the ability to shift branches.
In practice, the conditions under which chaos appears may become weaker because not only will equilibria shift location under slow variation of the mean depth, but so, and probably more importantly, will the corresponding domains of attraction. This is perhaps the reason for the apparently chaotic behavior found in Fig. 13 under a mild modulation of the forcing frequency. This figure shows the result of a numerical integration of (2) in phase space with a forcing of type (35). Both results of direct integration (Fig. 13a), as well as a sampling of these on the long, tidal timescale (Fig. 13b) are shown. The direct integration shows that the dominant resonant response on the short Helmholtz timescale vaccilates on the long timescale, between a small- and large-scale response. These small- and large-scale responses correspond to the amplified and choked regimes in the amplitude response diagram of Fig. 5 to which these solutions would tend in the absence of the tide (a = 0). Under modulation with a tide of amplitude a = 3ε2 this provokes a modulation in apparent frequency σ, of approximately 1.5 dimensionless units, which is not big enough to span the entire frequency range over which multiple equilibria exist. Both this figure, as well as the tidally subsampled version in Fig. 13b (with period 2π/ε2), however, testify about the irregular nature of the response, the latter figure in particular bearing some resemblance to the earlier Poincaré plots. In order to verify these qualitative ideas one may proceed with a more quantitative analysis. For the sake of brievity, this will however just be sketched here.
Several aspects of the numerical integration in Fig. 13 are still unrealistic. First, the response is dominated by the local resonance and thus the elevation shows a nearly periodic change on the fast timescale, instead of on the (long) tidal timescale. Second, the response is much too big, covering (almost) the entire basin depth. More work is therefore needed to bring the response closer to the ordering observed in nature, as for example, in Moldefjord (see Fig. 1), where Helmholtz response, O(0.1 m) ≪ tidal range, O(1 m) ≪ total depth, O(80 m), while currents on the fast Helmholtz timescale are comparable to tidal currents. It should be noticed that the correct ordering can partly be obtained quite simply by choosing model parameters such that all fixed points of the modulation equations are close to the origin. With this choice the homoclinic orbits, and therefore the actual state of the oscillator, are close to the origin of the phase space. Hence, corresponding elevations will be small compared to water depth. Notwithstanding the remaining issues that need clarification, it is believed that the mechanism described in this section may act as a “building block” for the explanation of the appearance of chaotic tides in basins with high Helmholtz frequency, particularly when the tide is extended with additional tidal components and harmonics.
From a practical point of view, any observed persistence of observed secondary oscillations might point at a non-meteorological origin. When their (average) period compares to estimates of the Helmholtz frequency, the model advanced here may be appropriate. Additional support may be obtained by plotting observed elevation (volume) against a strait current. Particularly, a stroboscopic sampling thereof (at the Helmholtz and tidal periods) may elucidate the structure of the underlying attractor. By comparison to predictions from the present model, an estimate of the most elusive factor, the damping, may perhaps be inferred.
6. Summary
Bays and estuaries may co-oscillate with the tides in the adjacent seas (Defant 1961). The extent to which they do generally depends on the geometry of the basin. Two competing mechanisms exist that mainly determine the final state of the coastal tides: the proximity of one of the more prolific tidal frequencies to basin resonances and the amount of friction. Depending on the balance between forcing and damping the coastal tide may be amplified or choked. This twofold nature of the response may actually occur for one and the same set of geometric and frictional parameters, as was shown in the case of the Helmholtz resonator for an almost-enclosed, relatively deep tidal basin with sloping bottom (see M). The slope in the bottom provokes a nonlinear restoring term (Green 1992) that, in turn, is responsible for the occurrence of multiple (stable) equilibria near the (linear) resonance frequency. When forced at the entrance by a singlefrequency tide (located within the resonance band), depending on the initial state, the tide within the basin may either have a small or large range that, after the decay of transients, is stationary.
Here we have shown that the presence of a second tidal frequency component, also within (or near to) the resonance band, may, in contrast, lead to the occurrence of a strange (chaotic) attractor, even with relatively strong damping. The basin tide may thus exhibit a perpetual change, manifested by unstable estimates of the tidal “constants” (amplitude and phase), when based on the analysis of a finite length time series, which may be of relevance to the observation by Gutiérrez et al. (1981) that some tidal components appear “unresolvable.” It may also be relevant to chaos in consecutive tidal maxima, observed in Venice Lagoon (Vittori 1992) or in coastal tides (Frison et al. 1999).
Surprisingly, there have been many observations of “irregular” secondary oscillations, accompanying the tide, which date back long before the notion of chaos appeared in the literature. Already in 1908, Honda et al. devoted an extensive study to ascertain the presence of both regular and irregular secondary oscillations in vertical elevation records in some 50, mostly Japanese, basins, usually based on observations covering just a few days. The nature of the basin response is typically, however, a superposition of the tide and a high-frequency oscillation, rendering the previous model, which predicts secondary oscillations on the tidal timescale, less applicable. Periods of these oscillations are observed to be in the range from several minutes up to several hours, depending on the size of the basin. Depending on the basin shape, this high-frequency oscillation can either be identified as the Helmholtz or quarter-wave resonance. The amplitude of the secondary oscillations in vertical elevation is, as the name suggests, observed to be an order of magnitude less than that of the (primary) tide. This may probably explain why, until fairly recently, the interest in secondary oscillations seems to have waned, and why the topic is held as a curiosity. A recent observation by Golmen et al. (1994), in a Norwegian fjord connected to the sea by a narrow strait, has, however, put the possible relevance of secondary oscillations into a new perspective by noticing that the associated secondary, irregular currents through the connecting strait reach a magnitude comparable to that of the tidal current. The reason for this is that, whereas the tide is present both outside and (with some delay) inside of the strait, the secondary elevation is present basically only within the fjord. The elevation differences, responsible for the associated currents, may therefore, at any time be of similar magnitude when comparing tidal and secondary oscillations. Hence, also the corresponding currents may be expected to be equally important. The same process may perhaps explain similar highly irregular current observations through lagoon entrances by Kjerfve and Knoppers (1991) and by Smith (1994), and the observed preferential enhancement of current harmonics (relative to that in the elevation field) by Seim and Sneed (1988). The relevance of this observation is that secondary oscillations may greatly alter the corresponding currents and therefore seriously modify the transfer of matter and dissolved substances (the “flushing” of the bay or fjord).
As remarked, these observations require a different kind of model in which tidal and resonance frequency—in our present model the Helmholtz frequency—are widely disparate. By adding a perturbative, but resonant, second forcing term to the primary tidal forcing it was shown that modulation equations are obtained that, to some extent, mimic those derived for the previous case with two tidal frequencies. In both cases, the modulation equations are driven at the long timescale. In the case of two nearby and resonant tidal frequencies this forcing stems from the beat (difference) frequency. In the case when the tidal frequency is much less than that of the resonant perturbative forcing term the tide itself is directly providing a modulation on the long timescale. The presence of a chaotic invariant set that could be ascertained analytically in these case, unfortunately does not guarantee that this set is also attractive. Therefore further support for the presence of a strange attractor was offered by (stroboscopic) Poincaré plots stemming from the numerical integration of the nonlinear Helmholtz resonator, when appropriately forced. The presence of a chaotic response may show up in a numerical simulation (in a basin with sloping shoreline) in much the same way as it reveals itself in nature: by irregularity in the elevation and, particularly, current fields (either within one tidal period, or in between subsequent tidal periods), by broadening of spectral peaks (with poorly resolved phases), and by showing a sensitive dependence on the initial tidal phase (neighboring orbits evolving to different states). The implication is that irregularity in the current and elevation fields is predictable as a phenomenon, but not its exact development. It requires further effort to quantify the resulting enhanced exchange between sea and basin.
Thus, despite the fact that, with the introduction of satellite-altimetry-derived tidal observations, tidal predictions have reached new unprecedented accuracies, it now seems that, at the same time, the tides may to some extent and in certain locations be intrinsically unpredictable, an unpredictability that reflects the nonlinear nature of the response of such locations.
Acknowledgments
Discussions with Marko Wilpshaar, and comments and suggestions by Jef Zimmerman, Huib de Swart, and Ferdinand Verhulst are greatly appreciated.
REFERENCES
Andersen, O. B., P. L. Woodworth, and R. A. Flather, 1995: Intercomparison of recent ocean tide models. J. Geophys. Res., 100 , 25261–25282.
Aref, H., 1984: Stirring by chaotic advection. J. Fluid Mech., 143 , 1–21.
Beigie, D., A. Leonard, and S. Wiggins, 1991: Chaotic transport in the homoclinic and heteroclinic tangle regions of quasiperiodically forced two-dimensional systems. Nonlinearity, 4 , 775–819.
Cartwright, D. E., 1977: Oceanic tides. Rep. Prog. Phys., 40 , 665.
Davies, A. M., 1986: A three-dimensional model of the northwest European continental shelf, with application to the M4 tide. J. Phys. Oceanogr., 16 , 797–813.
Defant, A., 1961: Physical Oceanography. Vol. II,. Pergamon, 598 pp.
Frison, T. W., H. D. I. Abarbanel, M. D. Earle, J. R. Schultz, and W. Scherer, 1999: Chaos and predictability in ocean water levels. J. Geophys. Res., 104 , 7935–7951.
Garrett, C., 1975: Tides in gulfs. Deep-Sea Res., 22 , 23–35.
Godin, G., 1991: The analysis of tides and currents. Tidal Hydrodynamics, B. B. Parker, Ed., Wiley and Sons, 675–709.
Golmen, L. G., J. Molvaer, and J. Magnusson, 1994: Sea level oscillations with super-tidal frequencies in a coastal embayment of western Norway. Cont. Shelf Res., 14 , 1439–1454.
Green, T., 1992: Liquid oscillations in a basin with varying surface area. Phys. Fluids A, 4 , 630–632.
Guckenheimer, J., and P. Holmes, 1983: Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields. Springer-Verlag, 459 pp.
Gutiérrez, A., F. Mosetti, and N. Purga, 1981: On the indetermination of the tidal harmonic constants. Nuovo Cimento, 4 , 563–575.
Honda, K., T. Terada, Y. Yoshida, and D. Isitani, 1908: Secondary undulations of oceanic tides. J. Coll. Sc. Imp. Univ. Tokyo, 24 , 1–113. and 95 plates.
Kjerfve, B., and B. A. Knoppers, 1991: Tidal choking in a coastal lagoon. Tidal Hydrodynamics, B. B. Parker, Ed., Wiley and Sons, 169–182.
Krümmel, O., 1911: Handbuch der Ozeanographie. Verlag von J. Engelhorns, 766 pp.
LeBlond, P. H., 1991: Tides and their interactions with other oceanographic phenomena in shallow water. Tidal Hydrodynamics, B. B. Parker, Ed., Wiley and Sons, 675–709.
Lynch, D. R., and F. E. Werner, 1991: Three-dimensional velocities from a finite-element model of English Channel/Southern Bight tides. Tidal Hydrodynamics, B. B. Parker, Ed., Wiley and Sons, 183–200.
Maas, L. R. M., 1997: On the nonlinear Helmholtz response of almost-enclosed tidal basins with a sloping bottom. J. Fluid Mech., 349 , 361–380.
Maas, L. R. M., . 1998: On an oscillator equation for tides in almost enclosed basins of non-uniform depth. Physics of Estuaries and Coastal Seas, J. Dronkers, and M. Scheffers, Eds., A. A. Balkema, 127–132.
Munk, W., and C. Wunsch, 1998: Abyssal recipes II: Energetics of tidal and wind mixing. Deep-Sea Res. I, 45 , 1977–2010.
Nakano, M., 1932: Preliminary note on the accumulation and dissipation of energy of the secondary oscillations in a bay. Proc. Phys.-Math. Soc. Japan, Ser. 3,, 14 , 44–56.
Nayfeh, A. H., and D. T. Mook, 1979: Nonlinear Oscillations. Wiley-Interscience, 704 pp.
Platzman, G., 1971: Ocean tides and related waves. Mathematical Problems in the Geophysical Sciences. Vol. 1, Geophysical Fluid Dynamics, W. H. Reid, Ed., American Mathematical Society, 239–291.
Prudnikov, A. P., Yu A. Brychkov, and O. I. Marichev, 1986: Integrals and Series. Vol. I, Gordon and Breach, 798 pp.
Pugh, D. T., 1987: Tides, Surges and Mean Sea-Level. John Wiley and Sons, 472 pp.
Ruelle, D., 1980: Strange attractors. Math. Intelligencer, 2 , 126–137. [Reprinted in Cvitanović, P., 1984: Universality in Chaos, Adam-Hilger, 37–48.].
Schwiderski, E. W., 1980: On charting global ocean tides. Rev. Geophys. Space Phys., 18 , 243–268.
Seim, H. E., and J. E. Sneed, 1988: Enhancement of semidiurnal tidal currents in the tidal inlets to Mississippi Sound. Hydrodynamics and Sediment Dynamics of Tidal Inlets, D. G. Aubrey and L. Weishar, Eds., Lecture Notes on Coastal and Estuarine Studies, Vol. 29, Springer-Verlag, 157–168.
Smith, N. P., 1994: Water, salt and heat balance of coastal lagoons. Coastal Lagoon Processes, B. Kjerfve, Ed., Elsevier Oceanographic Series, Vol. 60, 69–101.
van Ette, A. C. M., and H. J. Schoemaker, 1966: Harmonic analysis of tides. Essential features and disturbing influences. Hydrographic Letter, Vol. 1, Netherlands' Hydrographer Special Publication 2, 1–35.
Vittori, G., 1992: On the chaotic structure of tide elevation in the Lagoon of Venice. Proc. 23d Int. Conf. on Coastal Engineering, Venice, Italy, American Society of Civil Engineers, 1826–1839.
Wiggins, S., 1990: Introduction to Applied Nonlinear Dynamical Systems and Chaos. Springer-Verlag, 672 pp.
Yagasaki, K., 1990: Second-order averaging and chaos in quasiperiodically forced weakly nonlinear oscillators. Physica D, 44 , 445–458.
Yagasaki, K., . 1993: Chaotic motions near homoclinic manifolds and resonant tori in quasiperiodic perturbations of planar Hamiltonian systems. Physica D, 69 , 232–269.
Zimmerman, J. T. F., 1992: On the Lorentz linearization of a nonlinearly damped tidal Helmholtz oscillator. Proc. K. Ned. Akad. Wet., 95 , 127–145.
APPENDIX A
Integrating Eq. (21)
APPENDIX B
Computing the Melnikov Function
APPENDIX C
Evaluation of an Integral
Netherland Institute for Sea Research Publication Number 3288.