1. Introduction
Mesoscale oceanic variability is well known to be dominated by baroclinic instability (Gill et al. 1974), but some variability should be attributed to mesoscale atmospheric forcing (Orlanski and Polinsky 1983). As typical mesoscale atmospheric forcing, tropical cyclones (TCs) undoubtedly have the ability to produce mesoscale oceanic variability. This ability can be generally seen from the altimetry-based eddy changes caused by TCs (e.g., Lu et al. 2016; Ma et al. 2021; Shang et al. 2015; Sun et al. 2014; Walker et al. 2005, 2014). In addition to TC-induced mesoscale impacts, the large-scale ocean circulation in an ocean basin can also be modulated by the cumulative impacts of all the overlying TCs. For example, the summer cyclonic (anticyclonic) gyre in the northern (southern) South China Sea is enhanced by typhoons, i.e., TCs in the western North Pacific (WNP), while the basin-scale winter cyclonic gyre is enhanced (weakened) in the northern (southern) South China Sea (Ling et al. 2011; Wang et al. 2009). Zhang et al. (2020) showed that typhoons can accelerate the Kuroshio transport and further affect the extratropical ocean circulation and climate. However, these studies primarily estimated the integrated large-scale impacts of all the overlying typhoons. What dynamical processes dominate the TC-induced large-scale impacts still remains unclear.
TCs induce only two types of ocean dynamic responses in an open ocean, i.e., near-inertial and geostrophic responses (Geisler 1970; Greatbatch 1984; Nilsson 1995; Price 1983). Geostrophic response is induced by all TCs but near-inertial response only can be done for U > C, where U is the moving speed of a TC and C is the first baroclinic mode wave speed. Both geostrophic and near-inertial responses can affect the large-scale and mesoscale ocean circulation but in different ways. For near-inertial response, the ocean circulation is modulated by the enhanced vertical mixing subject to the dissipated near-inertial internal waves (Alford et al. 2016; Wunsch and Ferrari 2004), whereas the preexisting oceanic environment is linearly superimposed by the geostrophic response (Lu et al. 2020, hereafter LWS20). Owing to the uncertain locations of the dissipated internal waves and the inability of the merged altimetry datasets to capture the near-inertial signals, the altimetry-based eddy changes caused by TCs should arise from the geostrophic response. The estimated large-scale impacts in Zhang et al. (2020) are based on the altimetry-based eddy changes, and thus may be the result of the geostrophic response. Likewise, the large-scale impacts estimated in Wang et al. (2009) and Ling et al. (2011) should also be made via the geostrophic response because 1) the methods used in the two studies, Sverdrup theory and a 1.5-layer reduced gravity model, that did not parameterize the vertical mixing, and 2) the impacts was regarded to be relevant to the positive wind stress curl of typhoons, compatible with the mechanism of the geostrophic response (LWS20).
The geostrophic response consists of the barotropic and baroclinic components (Geisler 1970). The barotropic component is negligible in a deep ocean (Ginis and Sutyrin 1995) and what we call the geostrophic response is usually its baroclinic component. In merged altimetry datasets, the initial pattern of the geostrophic response is an along-track sea surface height (SSH) trough (Geisler 1970; Ginis and Sutyrin 1995; LWS20). The theoretical SSH trough is symmetrical about the track (Geisler 1970) within ±100 km of the track. The strength of the SSH trough is generally comparable with the amplitudes of the underlying eddies (LWS20). Thus, after the TC passage an along-track 200-km-width SSH trough should emerge in altimetry observations. This is true in some cases, for example, the observed SSH trough displayed in Fig. 1 in LWS20. However, in many cases the along-track SSH troughs are unobservable.
Many reasons may be responsible for the unobservable SSH trough. In theoretical treatments (Geisler 1970; Ginis and Sutyrin 1995; LWS20) and oceanic general circulation model (OGCM) experiments (Shay and Chang 1997; Shay et al. 1990), the ocean stratifications are usually assumed to be horizontally uniform so that the along-track SSH trough is easily seen from the theoretical and simulated results. The preexisting mesoscale variability in a real ocean may mask the superimposed SSH trough, enabling it to be unobservable. In addition, the altimetry-based SSH field is merged from along-satellite-track observations by optimal interpolation (Ducet et al. 2000; Pujol et al. 2016). The rapidly generated geostrophic response (LWS20; Price et al. 1994) cannot be portrayed by the interpolation of sparse along-satellite-track data. Thus, the merged SSH field near a TC passage must exist artificial smoothness (Lu et al. 2023). The artificial smoothness mitigates the strength of the SSH trough and makes it unobservable. Although unobservable in some cases, any observed long SSH trough can already prove the ability of the TC to impact the large-scale ocean circulation, thus being a prominent large-scale signal. As seen from Fig. 1 in LWS20, the oceanic circulation over the longitude–latitude range of 10° × 15° in 1–2 months is clearly regulated by the SSH trough induced by Typhoon Lupit (2003). Note that LWS20 did not mention the large-scale ocean impacts and only used this figure to stress the existence and importance of the geostrophic response.
This study has two purposes: 1) to present that the along-track jet and along-track SSH rises are another two observable signals for the large-scale impacts and 2) to demonstrate the large-scale impacts of nine typhoons (2015) on the subtropical gyre in the WNP. The mechanisms for the large-scale impacts are the same for global TCs; however, we only discuss the cases in the Northern Hemisphere. Since a rich-eddy zone exists over 21°–26°N, 127°–170°E dominated by baroclinic instability (Lin et al. 2008; Qiu 1999; Roemmich and Gilson 2001), this study primarily focuses on the large-scale impacts in the WNP below 20°N to avoid the mask of the preexisting eddies. Section 2 presents the theoretical background. In section 3, we demonstrate the large-scale impacts by numerical experiments. Section 4 introduces the observational data and method. Section 5 uses the observed large-scale signals for five typhoons (2015) to validate the theoretical and simulated results. Section 6 illustrates the more complicated large-scale impacts induced by some typhoons (2015) and discusses the existence of the large-scale impacts for the unobservable cases. Finally, we summarize the results in section 7.
2. Theoretical background
a. The SSH trough induced by a TC
The baroclinic geostrophic response on an f plane has been investigated in detail in LWS20. Although the TC moving direction and β effects play a negligible role in the generation of the geostrophic response (LWS20), they are vital for the post-TC evolution of the perturbed ocean. The features of the geostrophic response are summarized in the two-layer model presented in Geisler (1970).
As seen from Eq. (1), the positive curlzτ must lead to positive h and further the SSH (η) trough due to η = −hΔρ/ρ. For a constant forcing along the track, ∂h/∂X = 0 and ∂η/∂X = 0, so the baroclinic response manifests the along-track thermocline ridge and the along-track SSH trough. The geostrophic streamfunction and potential vorticity (PV) can be defined as ψTC = −g′h/f = gη/f and
As revealed in LWS20, the thermocline ridge or the SSH trough due to the upwelling is confined within |Y| ≤ Lh, sketched in Fig. 1, where the cross-track length scale Lh ≈ 100 km. The thermocline trough or the SSH rise associated with the downwelling is very weak. As can be seen from Fig. 1, a TC has the ability to perturb mesoscale eddies (large-scale ocean) in the cross-track (along-track) direction and the total track length (Ltrack) play a key role in the large-scale impacts. Generally, Ltrack ranges from several hundred to several thousand kilometers so that the large-scale impacts should not be ignored.
b. Quasigeostrophic motions after the TC passage
Three processes are involved in Eq. (3):
-
Quasigeostrophic adjustment driven by J(ψTC, qbasin) and J(ψbasin, qTC). Since ψbasin and qbasin is independent of the TC while ψTC and qTC have along-track patterns, the contours of ψTC(ψbasin) and qbasin(qTC) must not coincide so that J(ψTC, qbasin) ≠ 0 and J(ψbasin, qTC) ≠ 0. This suggests that the large-scale quasigeostrophic adjustment must occur after the TC passage.
-
Forced baroclinic Rossby waves. If the pre-TC ocean was presumed to be quiescent, Eq. (3) would be reduced towhich depicts forced baroclinic Rossby waves.
-
Free baroclinic Rossby waves. If the ocean is not passed by a TC, its motion is dominated by
This equation describes free baroclinic Rossby waves emanating from the east boundary of the ocean basin.
c. The linear baroclinic Rossby waves
After a TC passage, the evolution of the induced SSH trough should be closely related to its moving direction denoted by θ in Fig. 1. For θ = 0°, Ldkx ≈ Ld/Ltrack ≈ 0 and
3. The large-scale signals in OGCM experiments
a. Model configuration
To illustrate the evolutions of the SSH trough for different θ, we configured OGCM experiments on the β plane at the central latitude of 15°N using Hybrid Vertical Coordinate Ocean Model (HYCOM; Bleck 2002). The vertical mixing in the numerical experiments was chosen to be the K-profile scheme (Large et al. 1994). The advection–diffusion equations for temperature and salinity were treated by a variant (Drange and Bleck 1997) of the MPDATA scheme presented by Smolarkiewicz (1984). The horizontal resolution was 10 km, and the model was discretized using a 30-layer stretched vertical grid. The temperature and salinity were specified by Argo dataset ISAS20 (Gaillard et al. 2016) at 15°N, 152°E in March 2015. The model was run using the same grid in the (X, Y) coordinate but the results were displayed in the (x, y) coordinate. The Coriolis parameter on the β-plane experiments was set to f = f0 + β0y = f0 + β0 (−X sinθ + Y cosθ), where f0 = 3.77 × 10−5 s−1 and β0 = 2.21 × 10−11 m−1 s−1 for 15°N. To suppress the near-inertial waves, a radiation boundary condition was imposed along the four boundaries. In addition, we used a weak forcing with a small U to represent the real TC because near-inertial waves are hardly generated for U < C (Geisler 1970). We formulated the weak TC using the double exponential profile model in Willoughby et al. (2006) and made it move from X = 500 to −500 km at U = 6.94 km h−1. The maximum wind speed (Vmax) of this weak TC was specified to be 25.38 m s−1. The induced geostrophic response is virtually the same as that forced by the TC in LWS20 with Vmax = 35 m s−1 and U = 13.2 km h−1 since the geostrophic response is directly proportional to
Experiment descriptions.
We primarily demonstrate large-scale features based on the simulated SSH. Only the temperature anomaly, PV anomaly (PVA) and northward currents output from LIN_N were analyzed for the vertical structure of the evolved geostrophic response. The temperature anomaly and PVA were estimated relative to the initial values in quiescent ocean by using the methods in LWS20.
In addition to the above idealized experiments, the evolution of the SSH trough induced by Typhoon Higos (2015) with the real track and typhoon forcing was simulated to compare with altimetry observations. Most of the configuration for this case was the same as that in the idealized experiments. However, the initially quiescent ocean with the stratification at (13°N, 152°E) in February 2015 was forced by the typhoon with the real track and intensity. Another difference was that the model was configuration in longitude–latitude coordinate and the horizontal resolution was 0.08°.
b. The jet and the SSH rise emerged from linear evolutions
More clearly, the evolutions of the SSH trough depend heavily on the moving direction (Fig. 2). The SSH trough for θ = 0° virtually maintained the initially generated state (Figs. 2a,b) because it propagates as nondispersive long Rossby waves (section 2c). As θ = 45° or 90°, the initially 200-km-width SSH trough moved but dispersed significantly toward the west (Figs. 2c–f). Furthermore, after the short-term evolution a small-width along-track SSH rise emerged behind the respective trough (Figs. 2c,e). Subsequently, a secondary, weak along-track SSH trough was generated on the tail of the SSH rise during the longer evolution (Figs. 2d,f). The higher SSH gradient between the primary SSH trough and the SSH rise can be viewed as an along-track jet (Figs. 2c–e). In essence, these phenomena are the consequence of dispersive Rossby waves. As indicated in Eqs. (6) and (7), the lower wavenumber Rossby waves move westward faster than the higher wavenumber Rossby waves. Thus, for each y-section SSH trough, the longer wave components traveled faster so that the SSH gradients at the west (east) side were weakened (deepened). The deepened SSH gradient at the east side of the primary SSH trough forms an along-track jet, which is very similar to that of nonlinear self-advection occurring in the β-plane evolution of an isolated eddy (Early et al. 2011; McWilliams and Flierl 1979; Smith and O’Brien 1983). Since the higher wavenumber Rossby waves move more slowly, during the evolution they gradually seceded from the initial wave packet, i.e., the primary SSH trough, and formed Rossby wave wakes including the SSH rise and the secondary SSH trough. This process is the same as the emerged Rossby wave wakes behind the isolated eddy evolution (Early et al. 2011; McWilliams and Flierl 1979).
According to the nature of dispersive Rossby waves, the widths and strengths of the primary SSH trough, the SSH rise, and the jet for θ = 45° or 90° should vary with the time, represented in Fig. 3 by the features of LIN_NW and LIN_N in y = 0 section. With the dispersion of the primary SSH trough, its width (Ltrough) gradually grew (Fig. 3a). At the beginning, the width of the SSH rise (Lrise) was very large, ∼800 km, due to the contribution of downwelling (Fig. 3a). Then Lrise rapidly decreased in 10 days and then slowly decreased to an approximately steady value of 200–300 km. Owing to the wider initial perturbation for θ = 45°, Ltrough and Lrise in this case were always wider than those for θ = 90° (Fig. 3a). The strength of the primary SSH trough, denoted by the minimum SSH (
As classified in Camargo et al. (2007b), most of the moving directions for typhoons below 20°N belongs to θ ≤ 45° while the northward-moving typhoons mainly occurs above 20°N. These facts imply that the SSH trough in the WNP should usually disperse very slowly. In other words, the strengths and widths of the large-scale signals may vary slowly and not affect severely their observability in the WNP.
c. Contribution of downwelling
Figure 3a shows that the downwelling seemingly makes a contribution to the SSH rise. This contribution is evaluated by comparing the outputs of LIN_N and LIN_NnoDown. The simulated SSH trough looks like an athletics track (Fig. 4a) and the wider downwelling was located outside the narrower upwelling zone (Fig. 4b). However, the strength of the downwelling, with the maximum SSH of only ∼1.8 cm, was far smaller than that of the upwelling, consistent with Price (1981) and Jaimes and Shay (2015). The simulated evolution of the SSH trough by LIN_NnoDown (Figs. 4c,e) was almost the same as that by LIN_N (Figs. 2e,f). By contrast, in the sense of the linear dynamics, Figs. 4d and 4f approximately represent the evolution of the SSH rise in the downwelling zone, which was rather different than the evolution of the total SSH response in Figs. 2e and 2f. Moreover, the SSH contribution of downwelling was smaller than 1 cm during the evolution. Altogether, the Rossby wave wakes such as along-track SSH rise and jet primarily results from the dispersion of the primary SSH trough and the contribution of downwelling is ignorable.
d. Effects of nonlinearity
The linear evolutions in Fig. 2 are in fact dominated by the part ∂qTC/∂t + β∂ψTC/∂x = 0 in Eq. (4); thus, the nonlinear effects of J(ψTC, qTC) are omitted. Figure 5 shows the nonlinear evolutions to compare with Fig. 2. The results reveal that although Rossby wave dispersion is the main mechanism, the nonlinearity plays two roles in the evolutions. First, the nonlinear effects make the cyclonic ocean eddies (COEs) in the SSH troughs and the anticyclonic ocean eddies (AOEs) in the SSH rises to be formed because the local high concentrated PV can absorb the ambient like-sign PV to generate large coherent eddies by nonlinear interaction (McWilliams 1984). Consequently, the amplitudes of these COEs (AOEs) were slightly larger than the strengths of the corresponding SSH troughs (rises) in Fig. 2, while the strengths of the SSH troughs (rises) without eddies in Fig. 5 were obviously weaker than the corresponding strengths in Fig. 2. Under the action of the nonlinearity, the jet in Figs. 5c and 5d broke down finally and the COEs (AOEs) in the along-track SSH trough (rise) turned to be isolated. In this circumstance, the along-track patterns were not typically observable again. Second, similar to the self-advection in the evolution of an isolated eddy (Early et al. 2011; Smith and O’Brien 1983), the nonlinear advection associated with the along-track jet may inhibit Rossby wave dispersion. This inhabitation leads to the disappearance of the secondary SSH trough in Fig. 5d as compared with Fig. 2d. As explained in Early et al. (2011), Rossby wave dispersion drives the high SSH pattern toward the east but for θ = 45° the nonlinear advection linked with the jet does toward the northwest, so the nonlinear advection counteracts with Rossby wave dispersion. For θ = 90° the northward nonlinear advection has no effect on Rossby wave dispersion. For this reason, Rossby wave dispersion is very clear in Figs. 5e and 5f. By contrast, the interaction of the isolated COEs and AOEs dominates the evolution in Figs. 5c and 5d due to the inhibited Rossby wave dispersion.
e. Vertical structure
The vertical structure of the evolution is illustrated by temperature anomaly, υ and PVA in y = 0 section output from LIN_N (Fig. 6). A salient feature is that the TC-induced impacts are mainly restricted in the upper thermocline, which has been noted in LWS20. The initial vertical patterns (Figs. 6a,d,g) are almost the same as those in LWS20. The maximum temperature anomaly appeared at the depth of ∼150 m and the temperature response can reach 400 m (Fig. 6a). In addition, the geostrophic currents were located in 0–200-m depth (Fig. 6d), while the positive PVA in 100–200-m depth (Fig. 6g). With the evolution, the cold temperature anomaly and positive PVA in the primary SSH trough gradually decayed and the warm (cold) temperature anomaly and negative (positive) PVA in the SSH rise (the secondary trough) emerged (Figs. 6b,c,h,i). Accordingly, the strong geostrophic currents were also associated with the SSH rise and the secondary SSH trough (Figs. 6e,f). The TC-induced effects were still located in the upper thermocline although they slightly penetrated deeper in the vertical (Fig. 6). This shallower structure of the TC-induced impacts has a profound effect on the westward propagation speed of the observed large-scale signals, which will be discussed in section 5c.
4. Observational data and methods
a. Data
The best track dataset, International Best Track Archive for Climate Stewardship (IBTrACS v04r00; Knapp et al. 2010), was used to derive Vmax and to calculate U and Ltrack for the typhoons (Table 2). The typhoon intensity was denoted by the Saffir–Simpson scale categories from tropical storm (TS) to category 5. The initial SSH decreases are primarily located within ±Lh of the track (Fig. 1) so that a coverage named as COVER100 was calculated based on the track by setting Lh = 100 km (LWS20). The observed large-scale signals were identified from all-satellite-merged SSH dataset with 0.25° resolution downloaded from Copernicus Marine and Environment Monitoring Service (CMEMS). The along-satellite-track SSH data also from CMEMS were adopted to analyze the SSH changes along satellite tracks.
The mean SSH changes in COVER100 (
b. Methods
1) Observed large-scale SSH patterns
As seen in Fig. 1, a long-track TC has the ability to plow the preexisting SSH field. In fact, some typhoons in 2015 plowed the subtropical gyre in the WNP into several blocks of SSH troughs and SSH rises. To differentiate the block SSH trough (rise) with the along-track SSH trough (rise) induced by a single TC, we represented the former as acronym SSHT (SSHR). Sometimes the borders of the SSHTs and the SSHRs were the jet induced by a single typhoon. Using the eddy identification method in Chelton et al. (2011) and Xu et al. (2011), we identified and tracked these SSHTs, SSHRs, and borders to demonstrate the observed large-scale impacts on the subtropical gyre over 4°–20°N, 122°E–180°, the north boundary of which extended to 24°N in some cases. Furthermore, some COEs and AOEs were also identified to understand the relation between eddies and large-scale signals. We numbered all of the identified large-scale and mesoscale signals in chronological order of their occurrence. There were 11 typhoons that can induce the observed large-scale signals (Table 2). However, the signals induced by Typhoon Mekkhala were quickly masked by the ambient eddy field, and the track of Typhoon Higos was very short. Except for these two typhoons, below we demonstrate the large-scale impacts of the remaining nine typhoons on the WNP by tracking the identified SSHTs and SSHRs.
2) Mean SSH change in COVER100
In some cases, the large-scale signals cannot be identified from merged altimetry datasets. To quantify the TC-induced contributions, we estimated the mean SSH change (
Here the procedure to estimate
3) TC-induced SSH changes along satellite tracks
The observed large-scale signals in merged altimetry datasets may deviate from the SSH trough predicted by the linear theory in LWS20, particularly when the strong mesoscale variability preexists. This deviation may stem from the modification of the background geostrophic vorticity to upwelling (Jaimes and Shay 2009, 2015), the mitigation of the artificial smoothness (Lu et al. 2023), and the mask of background mesoscale variability. To clarify what factors cause the deviation, we estimated the along-satellite-track SSH changes, ΔηΑST. The tracks of the Jason-2 satellite with a 10-day repeated cycle were chosen here, and they should be approximately perpendicular to the TC track in order to investigate the symmetry of the SSH response. Then we estimated ΔηΑST along each of the chosen satellite tracks. To circumvent the possibly existing near-inertial signals, the post-TC track times were chosen to 5 ≤ t ≤ 8 days and thus the chosen pre-TC track times were −5 ≤ t ≤ −2 days. We further interpolated the merged dataset into the chosen satellite tracks and compared the original and merged ΔηΑST. This method will be used in section 6a.
5. Observational evidence for the theoretical and simulated results
a. Typhoon Mekkhala, θ ≈ 0°
A weak typhoon Mekkhala appeared in January 2015, and moved almost westward (Fig. 8), so θ ≈ 0° for this case. The preexisting mesoscale variability is so low (Fig. 8a) that the along-track SSH trough induced by this weak typhoon with TS scale is identifiable (Figs. 8b,c). The SSH trough mainly occurred in the COVER100, confirming that the cross-track length scale is ∼100 km. As seen in Table 2, the induced large-scale impacts were considerable due to the larger Ltrack (1738 km) and the moderate
b. Typhoons Bavi and Mayasak, θ ≈ 20°
As pointed out in section 2c, once θ increases from 0°, Rossby wave dispersion should emerge in the evolution of the SSH trough. Owing to no SSH rise and no jet for θ ≈ 0° (Figs. 2a,b, 5a,b, and 8b,c), it is necessary to answer for what minimum θ the SSH rise and jet will emerge. The supplemental numerical experiment show that the SSH rise and jet appear for θ = 15° (Fig. S8 in the online supplemental material). This result is also validated by altimetry observation for typhoons Bavi and Mayasak in March 2015. These two typhoons made a long journey over the WNP (Figs. 9b,c); both of their moving directions are θ ≈ 20°.
Prior to Bavi, the eddy activities were very weak except the along-track SSH trough (SSHT1) induced by Typhoon Higos (Figs. 9a and 10a–d). Although all the intensities of Bavi were the Saffir–Simpson scale of TS, most of the SSH trough induced by Bavi was visible (Fig. 9b). The mesoscale variability underlying Maysak was slightly stronger than that below Bavi and initially Maysak moved overlying a part of the SSH trough (SSHT2) induced by Bavi (Fig. 9c). However, the higher intensities of Maysak also generated an observable SSH trough (Fig. 9c). With the evolution of the SSH trough induced by Bavi, an along-track SSH rise and an along-track jet emerged clearly and both of them tended to be strengthened (Fig. 9c). An along-track jet and an along-track SSH rise seems to be formed from the SSH trough induced by Maysak but they were not very clear (Fig. 9c). Compared with Figs. 2 and 5, the secondary SSH troughs induced by the two typhoons cannot be identified possibly because of their weak strengths and the existed background mesoscale variability. Consistent with the simulated results, the strengths of the primary SSH troughs induced by the two typhoons were gradually weakened due to Rossby wave dispersion. Consequently, the weakened SSH trough was easily obscured by the background eddy field thus being unobservable (Figs. 9b,c). By contrast, the observed lifetimes of the jet and the SSH rise are far longer than that of the SSH trough (Fig. 9c).
Interestingly, the observed results for typhoons Bavi and Maysak resemble the linear evolutions very much in Figs. 2c and 2d rather than the nonlinear evolutions in Figs. 5c and 5d. During the nonlinear evolutions (Figs. 5c,d), the nonlinearity should cause a series strong COEs (AOEs) in the SSH trough (rise), which cannot be found in the observed evolutions, seeing Figs. 9c and 13a below. In particular, two COEs, C3 and C4, directly generated by Maysak were rapidly disrupted by Rossby wave dispersion. Unlike the simulated evolutions, in theory the observed evolutions must exist the effects of the free Rossby waves [Eqs. (3)–(5)]. Similar to the dispersion of forced Rossby waves, the dispersions of the free Rossby waves may inhibit the nonlinearity in the evolutions of the SSH trough.
As seen from the observed along-track signals for typhoons Bavi and Maysak (Fig. 9), the SSH field in the WNP was clearly plowed into three parts: the large SSHT below the borders B1 and B2 (SSHTlarge) including SSHT1 and SSHT2, the SSHR above the border B1 (SSHR1) and the SSHR semiclosed by the border B2 (SSHR2).
c. Typhoon Higos, θ ≈ 40°
Owing to the low SSH variability prior to Typhoon Higos (Fig. 10a), an SSH trough with two COEs (C1 and C2) was clearly seen (Figs. 10b,c). The moving direction was θ ≈ 40°, so the observations are well comparable to the idealized experiments for θ = 45°. Similar to Figs. 2c and 2d, the along-track jet emerged from the evolution of the SSH trough is discerned in Figs. 10b–d. However, the SSH rise seems visible but not very clear (Fig. 10c), possibly because the nonlinearity inhibits the forced Rossby wave dispersion (Fig. 5c). By comparing Fig. 5c with Fig. 10d, a remarkable difference can be found in westward propagation of the jet. After 30-day propagation, the simulated jet was still near the track (Fig. 5c), while the observed jet far away the track (Fig. 10d). It was hypothesized that the emerging COEs make the simulated jet slow down. To test this hypothesis, EXP_Higos (Table 1) was implemented and the SSH trough with two COEs were reproduced well (Figs. 10e,f). As indicated in Lu et al. (2023), the SSH extreme within the TC-induced COE can be viewed as the strength of the TC-induced geostrophic response. The simulated COE extreme for C1 (∼−14 cm) was slightly lower than the observed value (∼−12 cm) and the simulated value for C2 was well consistent with that from altimetry observation (Fig. 11a), suggesting that the geostrophic response to the typhoon was well simulated. However, the westward propagation speeds of the COEs were obviously slower (Figs. 10g,h) than what was observed in Figs. 10c and 10d, also seeing Fig. 11b.
If the westward propagation speeds of forced and free Rossby waves are
Using
Observed and simulated westward propagation speeds. The
This inconsistency in the observed and theoretical westward propagation speeds may be because the evolution of the SSH trough in a real ocean exists the vertical interactions between the forced and free Rossby waves. As hypothesized in section 5b, at least one of the roles of the vertical interaction is that free Rossby waves inhibit the nonlinearity in the SSH trough as the forced Rossby waves. For the Higos case, the inhibited nonlinearity can be seen from the fact that the simulated SSH trough was broken down early (Figs. 10g,h) but the observed SSH trough was long-lived (Figs. 10c,d). Unfortunately, the linear theory of White (1977) and Qiu et al. (1997) fails to treat the vertical interactions because in the theory the forced and free Rossby waves were assumed to occur in the same isopycnal layer. Indeed, a theory is required to understand the role of free Rossby waves but is beyond the scope of this study. Note that the observed strengths (Fig. 11a) and width (Fig. 10b) of the SSH trough are well consistent with the theory in LWS20 and the simulated results in Figs. 10e–h. Thus, the free Rossby waves may only affect the evolution of the SSH trough, not affecting the features related to the generation of the SSH trough such as its strength and width.
d. Typhoon Champi, θ ≈ 90°
After its formation on 14 October 2015, Typhoon Champi first moved westward as a weak and fast-moving typhoon and then turned to the north as a strong and slow-moving typhoon on 17 October 2015 (Fig. 12a). The observed SSH response induced by the westward-moving Champi was insignificant, while a strong SSH trough with a strong COE (C7) embedded appeared after 17 October 2015 (Fig. 12b). The northward SSH trough was approximately 1000 km long, very suited for validating the simulated cases in for θ ≈ 90° (Figs. 2e,f and 5e,f). Consistent with the simulated results, an along-track jet (B4) and an SSH rise (SSHR4) was generated on the east side of the SSH trough (Figs. 12a,b). Owing to the higher latitude, the Rossby wave dispersion is very weak (section 2c). However, after the long-term dispersion, the strengths of the SSH trough and C7 were still weakened and obscured by the ambient eddy field (Fig. 12c). The jet and the SSH rise were more robust than the SSH trough, less affected by the ambient eddy field (Fig. 12c), in line with the observation for Typhoon Bavi (Fig. 9c). As shown in Figs. 2 and 3, the strengths of the SSH rise and the jet decay due to the emerged secondary SSH trough. This can be supported by the fact that the strengths of the SSH rise and the jet in Fig. 12c were not to increase in contrast with Fig. 12b. However, we cannot assert that the secondary SSH trough on the tail of the SSH rise was the result of the Rossby wave dispersion (Fig. 12c).
Before Typhoon Champi, SSHR3 and B3 were formed and strengthened by several typhoons in July–August, seeing Fig. 15 below. Champi plowed the previous SSHR3 and B3 to generate SSHR4, B4, and C7 (Figs. 12a,b). Owing to the long-lived nature of geostrophic balance, the large-scale SSH patterns plowed by the previously observed large-scale signals moved westward until they encountered the Kuroshio. Most of them lasted until at least 31 December 2015; in particular, SSHR4 and B4 persisted until 15 February 2016 (Figs. 12c,d). Subsequently, SSHR4 degraded to an AOE, continuously moved westward and finally merged with the ambient AOE in mid-July 2016 (Fig. S9). It is clear that the impacts of Champi lasted about 9 months, from mid-October 2015 to mid-July 2016 (Fig. 12 and Fig. S9). The long-lived impacts mean that even if the local ocean is not passed over currently by any TC, the upper dynamics still can be affected by previous TCs. Although the TC-induced large-scale impacts should persist for long, the lifetime of the impacts induced by a specific typhoon is uncertain because it is determined by the distance between the impacts and west boundary currents and the interactions between the impacts and the ambient large-scale currents.
6. Complicated large-scale impacts induced by typhoons (2015) on subtropical gyre
In section 5, the observed large-scale impacts for five typhoon cases provide the strong evidence for the theoretical and simulated results because ψbasin and qbasin in Eq. (2) relative to ψTC and qTC are ignorable in these cases. In a real ocean, such observations are not frequently captured, and the observed large-scale impacts are usually very complicated. In this section, we illustrate the more complicated observations for several cases (sections 6a and 6b). In fact, even for the unobservable cases, the TC-induced large-scale impacts may also exist by quantifying the SSH changes in COVER100 (section 6c).
a. The along-track jet and the along-track SSH rise strengthened by Typhoon Dolphin
Unlike Figs. 2c and 2d, the along-track jet induced by Typhoon Dolphin seems not to emerge from the SSH trough (Fig. 13). Dolphin actually moved overlying the along-track jet (B1) induced by Typhoon Bavi. Before the arrival of Typhoon Dolphin, the jet may have been strengthening with the evolution, so it remains unclear whether the strengthened jet in Fig. 13b is closely related to Dolphin. To clarify this issue, the SSH (ηΑST) along three satellite tracks (L1–L3 in Fig. 13b) and their changes (ΔηΑST) are given in Fig. 14. The results show that Dolphin caused appreciable SSH decreases along the three tracks. Considering the existed errors of the TC track in best track datasets and the existed near-inertial signals in the post-TC original along-satellite-track data, both the original and merged SSH decreases along L1–L3 are thought to be approximately symmetrical about the track (Figs. 14d–f). The superposition of these approximately symmetrical SSH decreases to the pre-TC SSH did not lead to the symmetrical post-TC SSH (i.e., SSH trough) owing to the preexisting large cross-TC-track SSH gradients (Figs. 14a–c). Nevertheless, the contributions of Typhoon Dolphin strengthened the preexisting SSH gradients, particularly predominant in Figs. 14b and 14c. Thus, the large-scale impacts of Dolphin still existed in the forms of the strengthen B1 and the SSH rise on the right side of B1 (SSHR1). In addition, Dolphin further plowed the SSH field in 16°–20°N (Fig. 13b).
b. The combined large-scale impacts induced by multiple typhoons
In July 2015, two sequential typhoons, Chan-hom and Nangka, which mainly moved over the previous SSHR1 (Fig. 15a), caused significant sea surface cooling and mixed layer response (Wu and Li 2018). Over SSHR1, the intensities of Chan-hom were so weak that in this stage the along-track SSH trough induced by Chan-hom was not clearly observed (Fig. 15b). By contrast, although the along-track SSH trough induced by Typhoon Nangka was not identifiable over SSHR1, the associated along-track jet (B3) is roughly visible (Figs. 15a,b). Typhoons Chan-hom and Nangka passed successively over a preexisting AOE (A1), causing it to be disrupted and a large SSHT (SSHT3) to be generated (Figs. 15a,b). The emerged SSHT3 means that the observable impacts of Chan-hom and Nangka can extend to 24°N. The previous SSHR1 was plowed by Nangka into two parts: the smaller part was still termed as SSHR1 and the larger part was named as SSHR3 (Fig. 15b).
The observable signals can be found for three typhoons in August 2015 (Figs. 15b,c). The tracks of two sequential typhoons, Soudelor and Goni, are approximately parallel to those of Chan-hom and Nangka. Consequently, SSHT3 and a preexisting COE (C5) were significantly enlarged and enhanced (Figs. 15b,c), and a super COE (C6) was generated by triggering baroclinic instability. In addition, the strengths of both SSHR3 and the jet B3 were also enhanced. Only a part of the along-track SSH trough (SSHT4) induced by Typhoon Atsani is observable (Figs. 15b,c). However, the SSH rise and the jet above SSHT4 were connected to SSHR3 and B3, respectively. These connections also altered the large-scale SSH field significantly. Affected by the downwellings of Soudelor and Goni, both SSHR1 and SSHR2 were squeezed and tended to merge (Fig. 15c). SSHTlarge was not directly acted on by the five typhoons in July–August 2015, but it was also did indirectly by interacting continuously with SSHR1, SSHR2, and C6 (Figs. 15a–c).
c. Unobservable large-scale impacts
If none of the observable large-scale signals exist after the TC passage, do the large-scale impacts exist? As seen from
The TC-induced contributions of some typhoons may actually be underestimated by
The large-scale signals of the seven typhoons in Table 2 cannot be observed because of the preexisting strong mesoscale variability (Figs. S1–S7). Except for Typhoon Noul, the estimated
The group-mean
7. Conclusions
Previous studies have revealed that the large-scale ocean circulation in an ocean basin can be modulated by the cumulative impacts of all the overlying TCs. However, the underlying mechanism remains unrevealed. In this study, the mechanisms and the ability of a single TC to impact the large-scale ocean has been presented. The impacts are made via the geostrophic response, the observable signals for which may manifest as some of along-track SSH trough, along-track jet, and along-track SSH rise. These observed along-track signals suggest that the large-scale impacts are dominated by the total track length. Each of the long-track TCs has the ability to plow the preexisting SSH field significantly.
Shortly after the passage of a TC, the 200-km-width along-track SSH trough can always be viewed as the observable signal. However, if the moving direction of the TC has the northward component, an along-track jet and an along-track SSH rise will emerge due to Rossby wave dispersion of the SSH trough. By contrast, the SSH trough induced by a westward-moving TC approximately remains its initial state for long. These TC-induced large-scale impacts closely related to the moving directions are confirmed by idealized OGCM experiments and the observed signals induced by five typhoons (2015). In fact, the emerged jet and SSH rise imply that the initial SSH trough is widened and weakened by Rossby wave dispersion. Consequently, the SSH trough is easily masked by the ambient eddy field and thus the along-track jet and the SSH rise become the more identifiable signals. In some cases, the passage of the sequent typhoons over the large-scale signals formed by the previous typhoons makes the impacts of multiple typhoons more complicated to identify. However, even if the large-scale signals induced by a TC cannot be observed, the estimated mean SSH decreases in the COVER100 show that in most of such cases the large-scale impacts may still exist but merely cannot be seen intuitively.
In 2015, the large-scale signals of 11 long-track typhoons were observable. Except for Typhoons Mekkhala and Higos, each of the other 9 typhoons plow the SSH field over 4°–20°N, 122°E–180°) into several large SSHTs and SSHRs. These large, long-lived SSHTs and SSHRs can be plowed again by the subsequent typhoons, causing very complicated changes in the subtropical gyre. More importantly, the long-lived impacts imply that even if the local ocean is not passed over currently by any TC, the upper dynamics still can be affected by previous TCs via the geostrophic response.
Although only demonstrated by the cases in 2015, our results provide compelling evidence for the TC-induced large-scale impacts on subtropical gyre. The large-scale impacts are different each year because there is interannual variability in the TC tracks. In the WNP, the typhoon tracks can be classified into many clusters and the interannual variability of each cluster is related to large-scale atmospheric circulation and air–sea interaction such as an El Niño event (Camargo et al. 2007a,b). The particularly notable large-scale impacts in 2015 reported here may be closely related to an eastern Pacific El Niño event (Guo and Tan 2021). Despite this, at least 1–2 typhoons each year can induce approximately 1000-km-length observable large-scale signals (figures not given), suggesting that the large-scale impacts should be considered in future studies on upper-ocean dynamics in the WNP. This study paves a new way for understanding the large-scale impacts of a TC on upper-ocean dynamics. Undoubtedly, many questions remain to be answered. Specifically, the conditions regulating the observable and unobservable large-scale impacts need to be figured out. However, this question cannot be solved merely by altimetry observations but by in situ temperature–salinity observations because the observability depends on the vertical structure of the background field. In addition, a theory should be presented to explain the role of basin-scale free Rossby waves in the generation and persistence of the along-track jet and the along-track SSH rise.
Acknowledgments.
ZML is supported by the National Key Research and Development Program of China (2022YFF0801403), the National Natural Science Foundation of China (42276018), Independent Research Project Program of State Key Laboratory of Tropical Oceanography (LTOZZ2004). GHW is supported by the National Key Research and Development Program of China (2019YFC1510100) and the National Natural Science Foundation of China (41976003, 42030405). XDS is supported by the innovation project of Innovative Academy of Marine Information Technology, Chinese Academy of Sciences (CXBS202101). The numerical simulation is supported by the High Performance Computing Division in the South China Sea Institute of Oceanology. We gratefully acknowledge two anonymous reviewers for their constructive comments and suggestions.
Data availability statement.
The merged and along-track datasets are obtained from CMEMS with product identifiers of SEALEVEL_GLO_PHY_L4_REP_OBSERVATIONS_008_047 and SEALEVEL_GLO_PHY_L3_REP_OBSERVATIONS_008_062, respectively. The best track dataset (IBTrACS v04r00) is obtained from the best track datasets from https://www.ncei.noaa.gov/products/international-best-track-archive.
REFERENCES
Alford, M. H., J. A. MacKinnon, H. L. Simmons, and J. D. Nash, 2016: Near-inertial internal gravity waves in the ocean. Annu. Rev. Mar. Sci., 8, 95–123, https://doi.org/10.1146/annurev-marine-010814-015746.
Bleck, R., 2002: An oceanic general circulation model framed in hybrid isopycnic-Cartesian coordinates. Ocean Modell., 4, 55–88, https://doi.org/10.1016/S1463-5003(01)00012-9.
Camargo, S. J., A. W. Robertson, S. J. Gaffney, P. Smyth, and M. Ghil, 2007a: Cluster analysis of typhoon tracks. Part I: General properties. J. Climate, 20, 3635–3653, https://doi.org/10.1175/JCLI4188.1.
Camargo, S. J., A. W. Robertson, S. J. Gaffney, P. Smyth, and M. Ghil, 2007b: Cluster analysis of typhoon tracks. Part II: Large-scale circulation and ENSO. J. Climate, 20, 3654–3676, https://doi.org/10.1175/JCLI4203.1.
Chelton, D. B., and M. G. Schlax, 1996: Global observations of oceanic Rossby waves. Science, 272, 234–238, https://doi.org/10.1126/science.272.5259.234.
Chelton, D. B., R. A. deSzoeke, M. G. Schlax, K. El Naggar, and N. Siwertz, 1998: Geographical variability of the first baroclinic Rossby radius of deformation. J. Phys. Oceanogr., 28, 433–460, https://doi.org/10.1175/1520-0485(1998)028<0433:GVOTFB>2.0.CO;2.
Chelton, D. B., M. G. Schlax, R. M. Samelson, and R. A. de Szoeke, 2007: Global observations of large oceanic eddies. Geophys. Res. Lett., 34, L15606, https://doi.org/10.1029/2007GL030812.
Chelton, D. B., M. G. Schlax, and R. M. Samelson, 2011: Global observations of nonlinear mesoscale eddies. Prog. Oceanogr., 91, 167–216, https://doi.org/10.1016/j.pocean.2011.01.002.
Drange, H., and R. Bleck, 1997: Multidimensional forward-in-time and upstream-in-space-based differencing for fluids. Mon. Wea. Rev., 125, 616–630, https://doi.org/10.1175/1520-0493(1997)125<0616:MFITAU>2.0.CO;2.
Ducet, N., P. Y. Le Traon, and G. Reverdin, 2000: Global high-resolution mapping of ocean circulation from TOPEX/Poseidon and ERS-1 and -2. J. Geophys. Res., 105, 19 477–19 498, https://doi.org/10.1029/2000JC900063.
Early, J. J., R. Samelson, and D. B. Chelton, 2011: The evolution and propagation of quasigeostrophic ocean eddies. J. Phys. Oceanogr., 41, 1535–1555, https://doi.org/10.1175/2011JPO4601.1.
Gaillard, F., T. Reynaud, V. Thierry, N. Kolodziejczyk, and K. von Schuckmann, 2016: In situ–based reanalysis of the global ocean temperature and salinity with ISAS: Variability of the heat content and steric height. J. Climate, 29, 1305–1323, https://doi.org/10.1175/JCLI-D-15-0028.1.
Geisler, J. E., 1970: Linear theory of the response of a two layer ocean to a moving hurricane. Geophys. Fluid Dyn., 1, 249–272, https://doi.org/10.1080/03091927009365774.
Gill, A. E., 1982: Atmosphere–Ocean Dynamics. International Geophysics Series, Vol. 30, Academic Press, 662 pp.
Gill, A. E., J. S. A. Green, and A. J. Simmons, 1974: Energy partition in the large-scale ocean circulation and the production of mid-ocean eddies. Deep-Sea Res. Oceanogr. Abstr., 21, 499–528, https://doi.org/10.1016/0011-7471(74)90010-2.
Ginis, I., and G. Sutyrin, 1995: Hurricane-generated depth-averaged currents and sea surface elevation. J. Phys. Oceanogr., 25, 1218–1242, https://doi.org/10.1175/1520-0485(1995)025<1218:HGDACA>2.0.CO;2.
Greatbatch, R. J., 1984: On the response of the ocean to a moving storm: Parameters and scales. J. Phys. Oceanogr., 14, 59–78, https://doi.org/10.1175/1520-0485(1984)014<0059:OTROTO>2.0.CO;2.
Guo, Y.-P., and Z.-M. Tan, 2021: Influence of different ENSO types on tropical cyclone rapid intensification over the western North Pacific. J. Geophys. Res. Atmos., 126, e2020JD033059, https://doi.org/10.1029/2020JD033059.
Huang, R. X., and X. Jin, 2002: Sea surface elevation and bottom pressure anomalies due to thermohaline forcing. Part I: Isolated perturbations. J. Phys. Oceanogr., 32, 2131–2150, https://doi.org/10.1175/1520-0485(2002)032<2131:SSEABP>2.0.CO;2.
Jaimes, B., and L. K. Shay, 2009: Mixed layer cooling in mesoscale oceanic eddies during Hurricanes Katrina and Rita. Mon. Wea. Rev., 137, 4188–4207, https://doi.org/10.1175/2009MWR2849.1.
Jaimes, B., and L. K. Shay, 2015: Enhanced wind-driven downwelling flow in warm oceanic eddy features during the intensification of Tropical Cyclone Isaac (2012): Observations and theory. J. Phys. Oceanogr., 45, 1667–1689, https://doi.org/10.1175/JPO-D-14-0176.1.
Knapp, K. R., M. C. Kruk, D. H. Levinson, H. J. Diamond, and C. J. Neumann, 2010: The International Best Track Archive for Climate Stewardship (IBTrACS): Unifying tropical cyclone best track data. Bull. Amer. Meteor. Soc., 91, 363–376, https://doi.org/10.1175/2009BAMS2755.1.
Large, W. G., J. C. McWilliams, and S. C. Doney, 1994: Oceanic vertical mixing: A review and a model with a nonlocal boundary layer parameterization. Rev. Geophys., 32, 363–403, https://doi.org/10.1029/94RG01872.
Lin, I.-I., C.-C. Wu, I.-F. Pun, and D.-S. Ko, 2008: Upper-ocean thermal structure and the western North Pacific category 5 typhoons. Part I: Ocean features and the category 5 typhoons’ intensification. Mon. Wea. Rev., 136, 3288–3306, https://doi.org/10.1175/2008MWR2277.1.
Ling, Z., G. Wang, C. Wang, and Z.-S. Fan, 2011: Different effects of tropical cyclones generated in the South China Sea and the northwest Pacific on the summer South China Sea circulation. J. Oceanogr., 67, 347–355, https://doi.org/10.1007/s10872-011-0044-1.
Lu, Z., G. Wang, and X. Shang, 2016: Response of a preexisting cyclonic ocean eddy to a typhoon. J. Phys. Oceanogr., 46, 2403–2410, https://doi.org/10.1175/JPO-D-16-0040.1.
Lu, Z., G. Wang, and X. Shang, 2020: Strength and spatial structure of the perturbation induced by a tropical cyclone to the underlying eddies. J. Geophys. Res. Oceans, 125, e2020JC016097, https://doi.org/10.1029/2020JC016097.
Lu, Z., G. Wang, X. Shang, and X. Xie, 2023: Uncertainties in altimetry observations of eddy changes induced by tropical cyclones. J. Phys. Oceanogr., 53, 113–129, https://doi.org/10.1175/JPO-D-22-0115.1.
Ma, Z., Z. Zhang, J. Fei, and H. Wang, 2021: Imprints of tropical cyclones on structural characteristics of mesoscale oceanic eddies over the western North Pacific. Geophys. Res. Lett., 48, e2021GL092601, https://doi.org/10.1029/2021GL092601.
McWilliams, J. C., 1984: The emergence of isolated coherent vortices in turbulent flow. J. Fluid Mech., 146, 21–43, https://doi.org/10.1017/S0022112084001750.
McWilliams, J. C., and G. R. Flierl, 1979: On the evolution of isolated, nonlinear vortices. J. Phys. Oceanogr., 9, 1155–1182, https://doi.org/10.1175/1520-0485(1979)009<1155:OTEOIN>2.0.CO;2.
Nilsson, J., 1995: Energy flux from traveling hurricanes to the oceanic internal wave field. J. Phys. Oceanogr., 25, 558–573, https://doi.org/10.1175/1520-0485(1995)025<0558:EFFTHT>2.0.CO;2.
Orlanski, I., and L. J. Polinsky, 1983: Ocean response to mesoscale atmospheric forcing. Tellus, 35A, 296–323, https://doi.org/10.3402/tellusa.v35i4.11441.
Price, J. F., 1981: Upper ocean response to a hurricane. J. Phys. Oceanogr., 11, 153–175, https://doi.org/10.1175/1520-0485(1981)011<0153:UORTAH>2.0.CO;2.
Price, J. F., 1983: Internal wave wake of a moving storm. Part I. Scales, energy budget and observations. J. Phys. Oceanogr., 13, 949–965, https://doi.org/10.1175/1520-0485(1983)013<0949:IWWOAM>2.0.CO;2.
Price, J. F., 2022: 12.808 Supplemental Material, Topics in Fluid Dynamics: Dimensional Analysis, a Coriolis tutorial, and Lagrangian and Eulerian Representations, MIT Open Courseware, https://ocw.mit.edu/courses/res-12-001-topics-in-fluid-dynamics-spring-2022/pages/a-coriolis-tutorial/.
Price, J. F., T. B. Sanford, and G. Z. Forristall, 1994: Forced stage response to a moving hurricane. J. Phys. Oceanogr., 24, 233–260, https://doi.org/10.1175/1520-0485(1994)024<0233:FSRTAM>2.0.CO;2.
Pujol, M.-I., Y. Faugère, G. Taburet, S. Dupuy, C. Pelloquin, M. Ablain, and N. Picot, 2016: DUACS DT2014: The new multi-mission altimeter data set reprocessed over 20 years. Ocean Sci., 12, 1067–1090, https://doi.org/10.5194/os-12-1067-2016.
Qiu, B., 1999: Seasonal eddy field modulation of the North Pacific subtropical countercurrent: TOPEX/Poseidon observations and theory. J. Phys. Oceanogr., 29, 2471–2486, https://doi.org/10.1175/1520-0485(1999)029<2471:SEFMOT>2.0.CO;2.
Qiu, B., W. F. Miao, and P. Muller, 1997: Propagation and decay of forced and free baroclinic Rossby waves in off equatorial oceans. J. Phys. Oceanogr., 27, 2405–2417, https://doi.org/10.1175/1520-0485(1997)027<2405:PADOFA>2.0.CO;2.
Roemmich, D., and J. Gilson, 2001: Eddy transport of heat and thermocline waters in the North Pacific: A key to interannual/decadal climate variability? J. Phys. Oceanogr., 31, 675–687, https://doi.org/10.1175/1520-0485(2001)031<0675:ETOHAT>2.0.CO;2.
Shang, X.-d., H.-b. Zhu, G.-y. Chen, C. Xu, and Q. Yang, 2015: Research on cold core eddy change and phytoplankton bloom induced by typhoons: Case studies in the South China Sea. Adv. Meteor., 2015, 340432, https://doi.org/10.1155/2015/340432.
Shay, L. K., and S. W. Chang, 1997: Free surface effects on the near-inertial ocean current response to a hurricane: A revisit. J. Phys. Oceanogr., 27, 23–39, https://doi.org/10.1175/1520-0485(1997)027<0023:FSEOTN>2.0.CO;2.
Shay, L. K., S. W. Chang, and R. L. Elsberry, 1990: Free surface effects on the near-inertial ocean current response to a hurricane. J. Phys. Oceanogr., 20, 1405–1424, https://doi.org/10.1175/1520-0485(1990)020<1405:FSEOTN>2.0.CO;2.
Smith, D. C., and J. O’Brien, 1983: The interaction of a two-layer isolated mesoscale eddy with bottom topography. J. Phys. Oceanogr., 13, 1681–1697, https://doi.org/10.1175/1520-0485(1983)013<1681:TIOATL>2.0.CO;2.
Smolarkiewicz, P. K., 1984: A fully multidimensional positive definite advection transport algorithm with small implicit diffusion. J. Comput. Phys., 54, 325–362, https://doi.org/10.1016/0021-9991(84)90121-9.
Sun, L., Y.-X. Li, Y.-J. Yang, Q. Wu, X.-T. Chen, Q.-Y. Li, Y.-B. Li, and T. Xian, 2014: Effects of super typhoons on cyclonic ocean eddies in the western North Pacific: A satellite data-based evaluation between 2000 and 2008. J. Geophys. Res. Oceans, 119, 5585–5598, https://doi.org/10.1002/2013JC009575.
Walker, N. D., R. R. Leben, and S. Balasubramanian, 2005: Hurricane-forced upwelling and chlorophyll a enhancement within cold-core cyclones in the Gulf of Mexico. Geophys. Res. Lett., 32, L18610, https://doi.org/10.1029/2005GL023716.
Walker, N. D., R. R. Leben, C. T. Pilley, M. Shannon, D. C. Herndon, I.-F. Pun, I.-I. Lin, and C. L. Gentemann, 2014: Slow translation speed causes rapid collapse of northeast Pacific Hurricane Kenneth over cold core eddy. Geophys. Res. Lett., 41, 7595–7601, https://doi.org/10.1002/2014GL061584.
Wang, G., Z. Ling, and C. Wang, 2009: Influence of tropical cyclones on seasonal ocean circulation in the South China Sea. J. Geophys. Res., 114, C10022, https://doi.org/10.1029/2009JC005302.
White, W. B., 1977: Annual forcing of baroclinic long waves in the tropical North Pacific Ocean. J. Phys. Oceanogr., 7, 50–61, https://doi.org/10.1175/1520-0485(1977)007<0050:AFOBLW>2.0.CO;2.
Willoughby, H. E., R. W. R. Darling, and M. E. Rahn, 2006: Parametric representation of the primary hurricane vortex. Part II: A new family of sectionally continuous profiles. Mon. Wea. Rev., 134, 1102–1120, https://doi.org/10.1175/MWR3106.1.
Wu, R., and C. Li, 2018: Upper ocean response to the passage of two sequential typhoons. Deep-Sea Res. I, 132, 68–79, https://doi.org/10.1016/j.dsr.2017.12.006.
Wunsch, C., and R. Ferrari, 2004: Vertical mixing, energy, and the general circulation of the oceans. Annu. Rev. Fluid Mech., 36, 281–314, https://doi.org/10.1146/annurev.fluid.36.050802.122121.
Xu, C., X.-D. Shang, and R. X. Huang, 2011: Estimate of eddy energy generation/dissipation rate in the world ocean from altimetry data. Ocean Dyn., 61, 525–541, https://doi.org/10.1007/s10236-011-0377-8.
Zhang, Y., Z. Zhang, D. Chen, B. Qiu, and W. Wang, 2020: Strengthening of the Kuroshio current by intensifying tropical cyclones. Science, 368, 988–993, https://doi.org/10.1126/science.aax5758.