Skip to main content
Log in

An empirical model of tropical ocean dynamics

  • Published:
Climate Dynamics Aims and scope Submit manuscript

Abstract

To extend the linear stochastically forced paradigm of tropical sea surface temperature (SST) variability to the subsurface ocean, a linear inverse model (LIM) is constructed from the simultaneous and 3-month lag covariances of observed 3-month running mean anomalies of SST, thermocline depth, and zonal wind stress. This LIM is then used to identify the empirically-determined linear dynamics with physical processes to gauge their relative importance to ENSO evolution. Optimal growth of SST anomalies over several months is triggered by both an initial SST anomaly and a central equatorial Pacific thermocline anomaly that propagates slowly eastward while leading the amplifying SST anomaly. The initial SST and thermocline anomalies each produce roughly half the SST amplification. If interactions between the sea surface and the thermocline are removed in the linear dynamical operator, the SST anomaly undergoes less optimal growth but is also more persistent, and its location shifts from the eastern to central Pacific. Optimal growth is also found to be essentially the result of two stable eigenmodes with similar structure but differing 2- and 4-year periods evolving from initial destructive to constructive interference. Variations among ENSO events could then be a consequence not of changing stability characteristics but of random excitation of these two eigenmodes, which represent different balances between surface and subsurface coupled dynamics. As found in previous studies, the impact of the additional variables on LIM SST forecasts is relatively small for short time scales. Over time intervals greater than about 9 months, however, the additional variables both significantly enhance forecast skill and predict lag covariances and associated power spectra whose closer agreement with observations enhances the validation of the linear model. Moreover, a secondary type of optimal growth exists that is not present in a LIM constructed from SST alone, in which initial SST anomalies in the southwest tropical Pacific and Indian ocean play a larger role than on shorter time scales, apparently driving sustained off-equatorial wind stress anomalies in the eastern Pacific that result in a more persistent equatorial thermocline anomaly and a more protracted (and predictable) ENSO event.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

References

  • Alexander MA, Matrosova L, Penland C, Scott JD, Chang P (2008) Forecasting Pacific SSTs: linear inverse model predictions of the PDO. J Clim 21:385–402

    Article  Google Scholar 

  • An SI, Jin FF (2001) Collective role of thermocline and zonal advective feedbacks in the ENSO mode. J Clim 14:3421–3432

    Article  Google Scholar 

  • Anderson BT (2007) Intraseasonal atmospheric variability in the extratropics and its relation to the onset of tropical Pacific sea surface temperature anomalies. J Clim 20:926–936

    Article  Google Scholar 

  • Ashok K, Behere SK, Rao SA, Weng H, Yamagata T (2007) El Nino Modoki and its possible teleconnection. J Geophys Res 112. doi:10.1029/2006JC003798

  • Battisti DS, Hirst AC (1989) Interannual variability in a tropical atmosphere-ocean model: influence of the basic state, ocean geometry, and nonlinearity. J Atmos Sci 46:1687–1712

    Article  Google Scholar 

  • Bejarano L, Jin FF (2008) Coexistence of equatorial coupled modes of ENSO. J Clim 21:3051–3067

    Article  Google Scholar 

  • Bjerknes J (1969) Atmospheric teleconnections from the equatorial Pacific. Mon Weather Rev 97:163–172

    Article  Google Scholar 

  • Boulanger JP, Menkes C (2001) The Trident Pacific model. Part 2: role of long equatorial wave reflection on sea surface temperature anomalies during the 1993–1998 TOPEX/POSEIFON period. Clim Dyn 17:175–186

    Article  Google Scholar 

  • Burgers G, van Oldenborgh GJ (2003) On the impact of local feedbacks in the central Pacific on the ENSO cycle. J Clim 16:2396–2407

    Article  Google Scholar 

  • Burgers G, Jin FF, van Oldenborgh GJ (2005) The simplest ENSO recharge oscillator. Geophys Res Lett 32:L13706. doi:10.1029/2005GL022951

    Article  Google Scholar 

  • Capotondi A, Alexander MA (2001) Rossby waves in the tropical North Pacific and their role in decadal thermocline variability. J Phys Oceanogr 31:3496–3515

    Article  Google Scholar 

  • Carton JA, Giese BS (2008) A reanalysis of ocean climate using Simple Ocean Data Assimilation (SODA). Mon Weather Rev 136:2999–3017

    Article  Google Scholar 

  • Chang P, Zhang L, Saravanan R, Vimont DJ, Chiang JCH, Ji L, Seidel H, Tippett MK (2007) Pacific meridional mode and El Niño-Southern Oscillation. Geophys Res Lett 34:L16608. doi:10.1029/2007GL030302

    Article  Google Scholar 

  • Chelton DB, Schlax MG (1996) Global observations of oceanic Rossby waves. Science 272:234–238

    Article  Google Scholar 

  • Chelton DB, Wentz FJ, Gentemann CL, de Szoeke RA, Schlax MG (2000) Satellite microwave SST observations of transequatorial tropical instability waves. Geophys Res Lett 27:1239–1242

    Article  Google Scholar 

  • Chen D, Cane MA, Kaplan A, Zebiak SE, Huang D (2004) Predictability of El Niño over the past 148 years. Nature 428:733–736

    Article  Google Scholar 

  • Clarke AJ, Van Gorder S (2003) Improving El Niño prediction using a space-time integration of Indo-Pacific winds and equatorial Pacific upper ocean heat content. Geophys Res Lett 30:1399. doi:10.1029/2002GL016673

    Article  Google Scholar 

  • Dijkstra HA, Neelin JD (1999) Coupled processes and the tropical climatology. Part III: instabilities of the fully coupled climatology. J Clim 12:1630–1643

    Article  Google Scholar 

  • Farrell B (1988) Optimal excitation of neutral Rossby waves. J Atmos Sci 45:163–172

    Article  Google Scholar 

  • Fedorov AV, Philander SG (2001) A stability analysis of tropical ocean-atmosphere interactions: Bridging measurements and theory for El Niño. J Clim 14:3086–3101

    Article  Google Scholar 

  • Flügel M, Chang P (1996) Impact of dynamical and stochastic processes on the predictability of ENSO. Geophys Res Lett 23:2089–2092

    Article  Google Scholar 

  • Frankignoul C, Hasselman K (1977) Stochastic climate models. Part II: application to sea-surface temperature variability and thermocline variability. Tellus 29:289–305

    Article  Google Scholar 

  • Guilyardi E, Wittenberg A, Fedorov A, Collins M, Wang C, Capotondi A, van Oldenborgh GJ, Stockdale T (2009) Understanding El Niño in ocean-atmosphere general circulation models: progress and challenges. Bull Am Meteor Soc 90:325–340

    Article  Google Scholar 

  • Hasselmann K (1976) Stochastic climate models. Part I. Theory. Tellus 28:474–485

    Article  Google Scholar 

  • Jiang N, Neelin JD, Ghil M (1995) Quasi-quadrennial and quasi-biennial variability in the equatorial Pacific. Clim Dyn 12:101–112

    Google Scholar 

  • Jin FF (1997) An equatorial ocean recharge paradigm for ENSO. Part I: conceptual model. J Atmos Sci 54:811–829

    Article  Google Scholar 

  • Jin FF, An SI (1999) Thermocline and zonal advective feedbacks within the equatorial ocean recharge oscillator model for ENSO. Geophys Res Lett 26:2989–2992

    Article  Google Scholar 

  • Jin FF, Neelin JD (1993) Modes of interannual tropical ocean-atmosphere interaction—a unified view. Part III: analytical results in fully coupled cases. J Atmos Sci 50:3523–3540

    Article  Google Scholar 

  • Jin FF, Neelin JD, Ghil M (1994) El Nino on the “Devils Staircase”: annual subharmonic steps to chaos. Science 264:710–713

    Article  Google Scholar 

  • Jin FF, Kim ST, Bejarano L (2006) A coupled-stability index for ENSO. Geophys Res Lett 33:L23708. doi:10.1029/2006GL027221

    Article  Google Scholar 

  • Jochum M, Cronin MF, Kessler WS, Shea D (2007) Observed horizontal temperature advection by tropical instability waves. Geophys Res Lett 34:L09604. doi:10.1029/2007GL029416

    Article  Google Scholar 

  • Johnson SD, Battisti DS, Sarachik ES (2000) Empirically derived Markov models and prediction of tropical pacific sea surface temperature anomalies. J Clim 13:3–17

    Article  Google Scholar 

  • Kallummal R, Kirtman B (2008) Validity of a linear stochastic view of ENSO in an ACGCM. J Atmos Sci 65:3860–3879. doi:10.1175/2008JAS2286.1

    Article  Google Scholar 

  • Kalnay E et al (1996) The NCEP/NCAR 40-Year reanalysis project. Bull Am Meteor Soc 77:437–471

    Article  Google Scholar 

  • Kang IS, An SI, Jin FF (2001) A systematic approximation of the SST anomaly equation for ENSO. J Met Soc Jpn 79:1–10

    Article  Google Scholar 

  • Kao HY, Yu JY (2009) Contrasting Eastern-Pacific and Central-Pacific types of ENSO. J Clim 22:615–632

    Google Scholar 

  • Kessler WS (2002) Is ENSO a cycle or a series of events? Geophys Res Lett 29:2125. doi:10:1029/2002GL015924

    Article  Google Scholar 

  • Kessler WS, McPhaden MJ (1995) The 1991–93 El Niño in the central Pacific. Deep-Sea Res II 42:295–334

    Article  Google Scholar 

  • Latif M, Graham NE (1992) How much predictive skill is contained in the thermal structure of an OGCM? J Phys Oceanogr 22:951–962

    Article  Google Scholar 

  • MacMynowski DG, Tziperman E (2008) Factors affecting ENSO’s period. J Atmos Sci 65:1570–1586

    Article  Google Scholar 

  • McPhaden MJ (1999) Genesis and evolution of the 1997–98 El Niño. Science 283:950–954

    Article  Google Scholar 

  • McPhaden MJ, Zhang X, Hendon HH, Wheeler MC (2006) Large scale dynamics and MJO forcing of ENSO variability. Geophy Res Lett 33:L16702. doi:10.1029/2006GL026786

    Article  Google Scholar 

  • Meinen CS, McPhaden MJ (2000) Observations of warm water volume changes in the equatorial pacific and their relationship to El Nino and La Nina. J Clim 13:3551–3559

    Article  Google Scholar 

  • Moore A, Kleeman R (1997) The singular vectors of a coupled-atmosphere model of ENSO. II: sensitivity studies and dynamical interpretation. Quart J R Meteor Soc 123:983–1006

    Article  Google Scholar 

  • Moore A, Kleeman R (1999) The non-normal nature of El Niño and intraseasonal variability. J Clim 12:2965–2982

    Article  Google Scholar 

  • Neelin JD (1991) The slow sea surface temperature mode and the fast-wave limit: analytic theory for tropical interannual oscillations and experiments in hybrid coupled models. J Atmos Sci 48:584–606

    Article  Google Scholar 

  • Neelin JD, Jin FF (1993) Modes of interannual tropical ocean-atmosphere interaction—a unified view. Part II: analytical results in the weak-coupling limit. J Atmos Sci 50:3504–3522

    Article  Google Scholar 

  • Newman M (2007) Interannual to decadal predictability of tropical and North Pacific sea surface temperatures. J Clim 20:2333–2356

    Article  Google Scholar 

  • Newman M, Sardeshmukh PD (2008) Tropical and stratospheric influences on extratropical short-term climate variability. J Clim 21:4326–4347

    Article  Google Scholar 

  • Newman M, Sardeshmukh PD, Winkler CR, Whitaker JS (2003) A study of subseasonal predictability. Mon Weather Rev 131:1715–1732

    Article  Google Scholar 

  • Newman M, Sardeshmukh PD, Penland C (2009) How important is air-sea coupling in ENSO and MJO evolution? J Clim 22:2958–2977

    Article  Google Scholar 

  • Ortiz-Bevía MJ (1997) Estimation of the cyclostationary dependence in geophysical data fields. J Geophys Res 102:13473–13486

    Article  Google Scholar 

  • Papanicolaou G, Kohle W (1974) Asymptotic theory of mixing stochastic ordinary differential equations. Commun Pure Appl Math 27:641–668

    Article  Google Scholar 

  • Penland C (1989) Random forcing and forecasting using principal oscillation pattern analysis. Mon Weather Rev 117:2165–2185

    Article  Google Scholar 

  • Penland C (1996) A stochastic model of IndoPacific sea surface temperature anomalies. Phys D 98:534–558

    Article  Google Scholar 

  • Penland C, Matrosova L (1994) A balance condition for stochastic numerical models with application to the El Niño-Southern Oscillation. J Clim 7:1352–1372

    Article  Google Scholar 

  • Penland C, Matrosova L (2006) Studies of El Niño and interdecadal variability in Tropical sea surface temperatures using a nonnormal filter. J Clim 19:5796–5815

    Article  Google Scholar 

  • Penland C, Sardeshmukh PD (1995) The optimal growth of tropical sea surface temperature anomalies. J Clim 8:1999–2024

    Article  Google Scholar 

  • Penland C, Flügel M, Chang P (2000) Identification of dynamical regimes in an intermediate coupled ocean-atmosphere model. J Clim 13:2105–2115

    Article  Google Scholar 

  • Picaut J, Ioualalen M, Menkes C, Delcroix T, McPhaden MJ (1996) Mechanism of the zonal displacements of the Pacific Warm Pool: Implications for ENSO. Science 274:1486–1489

    Article  Google Scholar 

  • Rayner NA, Parker DE, Horton EB, Folland CK, Alexander LV, Rowell DP, Kent EC, Kaplan A (2003) Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. J Geophys Res 108:4407. doi:10.1029/2002JD002670

    Article  Google Scholar 

  • Schopf PS, Suarez MJ (1988) Vacillations in a coupled ocean-atmosphere model. J Atmos Sci 45:549–566

    Article  Google Scholar 

  • Schopf PS, Anderson DLT, Smith R (1981) Beta-dispersion of low frequency Rossby waves. Dyn Atmos Oceans 5:187–214

    Article  Google Scholar 

  • Thompson CJ, Battisti DS (2001) A linear stochastic dynamical model of ENSO. Part II: analysis. J Clim 14:445–466

    Article  Google Scholar 

  • Vimont DJ, Wallace JM, Battisti DS (2003) The seasonal footprinting mechanism in the Pacific: implications for ENSO. J Clim 16:2668–2675

    Article  Google Scholar 

  • Wang W, McPhaden MJ (2000) The surface-layer heat balance in the equatorial Pacific ocean. Part II: interannual variability. J Phys Ocean 30:2989–3008

    Article  Google Scholar 

  • Wang X, Jin FF, Wang Y (2003) A tropical ocean recharge mechanism for climate variability. Part I: equatorial heat content changes induced by the off-equatorial wind. J Clim 16:3585–3598

    Article  Google Scholar 

  • White WB, Tourre YM, Barlow M, Dettinger M (2003) A delayed action oscillator shared by biennial, interannual, and decadal signals in the Pacific Basin. Geophys Res 108. doi:10.1029/2002JC001490

  • Winkler CR, Newman M, Sardeshmukh PD (2001) A linear model of wintertime low-frequency variability. Part I: formulation and forecast skill. J Clim 14:4474–4494

    Article  Google Scholar 

  • Wyrkti K (1985) Water displacements in the Pacific and the genesis of El Niño cycles. J Geophys Res 90:7129–7132

    Article  Google Scholar 

  • Xue Y, Leetmaa A, Ji M (2000) ENSO prediction with Markov models: the impact of sea level. J Clim 13:849–871

    Article  Google Scholar 

  • Zebiak SE, Cane MA (1987) A model El Niño-southern oscillation. Mon Weather Rev 115:2268–2278

    Article  Google Scholar 

  • Zelle H, Appeldoorn G, Burgers G, van Oldenborgh GJ (2004) The relationship between sea surface temperature and thermocline depth in the eastern equatorial Pacific. J Phys Ocean 34:643–655

    Article  Google Scholar 

  • Zhang Y, Wallace JM, Battisti DS (1997) ENSO-like interdecadal variability. J Clim 10:1004–1020

    Article  Google Scholar 

Download references

Acknowledgments

The authors thank Antonietta Capotondi, Cécile Penland, Amy Solomon, David Battisti and two anonymous reviewers for helpful comments. This work was partially supported by a grant from NOAA CLIVAR-Pacific.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Matthew Newman.

Appendix: Construction and validation of the LIM

Appendix: Construction and validation of the LIM

In any multidimensional statistically stationary system with components x i , one may define a time lag covariance matrix C(τ) with elements \( Cij\left( \tau \right) = \left\langle {x_{i} (t + \tau )x(t)} \right\rangle \), where angle brackets denote a long term average. In linear inverse modeling, one assumes that the system satisfies C(τ) = G(τ)C(0), where importantly G(τ) = exp(Lτ) and L is a constant matrix, which follows from (2). One then uses this relationship to estimate L from observational estimates of C(0) and C 0 ) at some lag τo. For the statistics of this system to be stationary, L must be dissipative, i.e its eigenvalues must have negative real parts. In a forecasting context, G(τ)x(t) represents the “best” forecast (in a least squares sense) of x(t + τ) given x(t). Note that unlike multiple linear regression, determination of G at one lag τo identically gives G at all other lags. Also, statistics of the noise forcing, not just the error covariance, are determined by LIM since the positive-definite noise covariance matrix \( {\mathbf{Q}} = \left\langle {\xi \xi^{T} } \right\rangle dt \) is determined from a Fluctuation-Dissipation relationship (3), given the observed C(0) and L.

A training lag of το = 3 months was used to determine L. The EOF truncations (Sect. 2) and training lag were chosen to maximize the LIM’s cross-validated forecast skill for leads up to 18 months, while avoiding the Nyquist problem for L (PS95) that inhibits analysis of interactions amongst the model variables. In no other respect do these choices qualitatively affect the points made in this paper. Estimates of L and of forecast skill were cross-validated by sub-sampling the data record by sequentially removing one five-year period, computing L for the remaining years, and then generating forecasts for the independent years. This procedure was repeated for the entire period. Forecast skill is then determined by comparing the local anomaly correlation between the cross-validated model predictions and gridded untruncated verifications.

Explicitly including both τx and Z 20 in the LIM state vector increases skill of 9 (Figs. 12a, b) and 18 (Fig. 12c, d) month T o forecasts. Skill improvement is mostly due to the inclusion of Z 20 rather than τx, confirming that the Z 20 data provides useful information. In addition, the LIM makes Z 20 forecasts whose skill is significantly better than persistence (not shown). The difference in SST skill between the ocean and SST-LIMs generally increases with forecast lead because the ocean LIM captures the slower Z 20 evolution. For example, a second set of “fixed-Z 20” ocean LIM forecasts in which the initial Z 20 anomaly is persisted throughout the forecast period (i.e., Z 20(t) = Z 20(0); not shown) has greatly reduced 18-month forecast skill.

Fig. 12
figure 12

Cross-validated forecast skill (1959–2000) of T O for forecast leads of 9 and 18 months, for the LIM (top) and the SST13-LIM (bottom)

We test the validity of the linear approximation with a “tau-test” (PS95). For example, since (2) implies that C(τ) = G(τ) C(0), the LIM should be able to reproduce observed lag-covariance statistics at much longer lags than the 3 month lag on which the LIM was trained (e.g., C(18) = [G(3)]6 C(0)). Figure 13 compares the observed and predicted lag-autocovariances of T O for lags of 9, 18, and 36 months, using both the ocean LIM and the SST13-LIM. The SST13-LIM does a reasonably good job capturing the main aspects of the lag-autocovariance pattern, but for lags less than about a year it tends to overestimate persistence especially along the equator and for longer lags it errs in the amplitude of the negative lag-autocovariance. The ocean LIM improves upon these deficiencies, as well as reproducing observed lag-covariance for Z 20 over the same lags (not shown).

Fig. 13
figure 13

Observed (top panels), LIM (middle panels) and SST13-LIM (bottom panels) T O lag-covariance. (left) 9-month lag-covariance; (center) 18-month lag-covariance; (right) 36-month lag-covariance. Note that the observed lag-covariances are based on the full (that is, not truncated in the EOF basis) gridded anomaly fields. Contour interval is 0.04 K2

A complementary test of linearity is to compare observed and LIM-predicted power spectra, by integrating (2) for 42,000 years using the method described in Penland and Matrosova (1994) and Newman et al. (2009) and then collecting statistics. The white noise forcing is determined from the noise covariance matrix Q determined as a residual from (3). Q should be positive-definite but determined this way it is only guaranteed to be symmetric. Ensuring positive-definiteness in the manner of Penland and Matrosova (1994), by rescaling the noise due to one small negative eigenvalue of Q that accounted for less than 0.25% of the trace of Q, resulted in an almost negligible impact on C(0). The resulting model “data” is separated into 1,000 42-year time series. The observed spectra and the ensemble mean of the model spectra for the three leading PCs of T O and Z 20 are shown in Figs. 14 and 15, respectively. The corresponding EOF pattern for each spectrum is shown in the inset panels. The gray shading shows the 95% confidence intervals of these spectra, estimated using the 1,000 model realizations.

Fig. 14
figure 14

Power spectra for the three leading SST (T O ) PCs (red lines), compared to that predicted by the LIM (blue lines). Gray shading represents the 95% confidence interval determined from a 1,000-member ensemble of 42 year LIM model runs (see text for further details). In these log(frequency) versus power times angular frequency (ω) plots, the area under any portion of the curve is equal to the variance within that frequency band. Note that displaying power times frequency slightly shifts the power spectral density peak centered at a period of 4.5 years to a variance peak centered at a period of 3.5 years. Insets in each panel show the corresponding EOF and the variance explained by that pattern

Fig. 15
figure 15

Same as Fig. 14 but for the three leading 20°C isotherm depth (Z 20) PCs. Note that there is no SST-LIM equivalent

The LIM reproduces the main features of the observed power spectra for the leading PCs of each variable (including τx, not shown). Obviously, the mean LIM spectra are much smoother than observed, due to the relatively few degrees of freedom in the truncated EOF space. On the other hand, the irregularity of the observed spectra is at least partly due to sampling, as indicated by the confidence intervals, which show how much variation in the spectra could occur simply from different realizations of noise.

Since C(0) and C(τ) have seasonal dependence, both Johnson et al. 2000 and Xue et al. 2000 suggested that L should also be considered to be seasonally-varying. They constructed “Markov models” in which they determined Gο) for each season (but not L) so that forecasts are made by an appropriate product of each G. On the other hand, PS95 and Penland (1996) argued that the observed seasonality of ENSO, including its phase locking, can be captured with a fixed L with seasonally varying Q. Newman et al. (2009) found that a tropical atmosphere-SST LIM constructed using weekly data had a poorer representation of coupled ENSO dynamics when segregated by season, although the internal subseasonal atmospheric dynamics were slightly improved. They also found pronounced seasonality of the noise, as did Penland (1996) and Chang et al. (2007). Similar to Penland, we found that the seasonal dependence of the tests above can be generally captured by assuming fixed L but with seasonally varying Q. As in Newman et al. (2009), we also found that seasonal LIMs did poorer in these tests compared to those using a fixed L.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Newman, M., Alexander, M.A. & Scott, J.D. An empirical model of tropical ocean dynamics. Clim Dyn 37, 1823–1841 (2011). https://doi.org/10.1007/s00382-011-1034-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00382-011-1034-0

Keywords

Navigation