This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Rigorous Derivation of Stochastic Conceptual Models for the El Niño-Southern Oscillation from a Spatially-Extended Dynamical System

Nan Chen Department of Mathematics, University of Wisconsin-Madison, 480 Lincoln Dr., Madison, WI 53706, USA Yinling Zhang [email protected] Department of Mathematics, University of Wisconsin-Madison, 480 Lincoln Dr., Madison, WI 53706, USA
Abstract

El Niño-Southern Oscillation (ENSO) is the most predominant interannual variability in the tropics, significantly impacting global weather and climate. In this paper, a framework of low-order conceptual models for the ENSO is systematically derived from a spatially-extended stochastic dynamical system with full mathematical rigor. The spatially-extended stochastic dynamical system has a linear, deterministic, and stable dynamical core. It also exploits a simple stochastic process with multiplicative noise to parameterize the intraseasonal wind burst activities. A principal component analysis based on the eigenvalue decomposition method is applied to provide a low-order conceptual model that succeeds in characterizing the large-scale dynamical and non-Gaussian statistical features of the eastern Pacific El Niño events. Despite the low dimensionality, the conceptual modeling framework contains outputs for all the atmosphere, ocean, and sea surface temperature components with detailed spatiotemporal patterns. This contrasts with many existing conceptual models focusing only on a small set of specified state variables. The stochastic versions of many state-of-the-art low-order models, such as the recharge-discharge and the delayed oscillators, become special cases within this framework. The rigorous derivation of such low-order models provides a unique way to connect models with different spatiotemporal complexities. The framework also facilitates understanding the instantaneous and memory effects of stochastic noise in contributing to the large-scale dynamics of the ENSO.

keywords:
Stochastic conceptual model, Eigenvalue decomposition, Spatially-extended stochastic dynamical systems, Large-scale ENSO features, Multiplicative noise, Stochastic discharge-recharge oscillator, Stochastic delayed oscillator
MSC:
[2010] 35R60, 37N10, 60H10 , 65F15

1 Introduction

El Niño-Southern Oscillation (ENSO) is the most predominant interannual variability in the tropics with significant impacts on the global weather and climate [1, 2, 3, 4]. El Niño and La Niña are opposing patterns, representing the warm and cool phases of a recurring climate pattern across the tropical Pacific. They form an irregular oscillator, shifting back and forth every three to seven years, leading to predictable shifts in sea surface temperature (SST) and disrupting the wind and rainfall patterns across the tropics.

A hierarchy of dynamical and statistical models with different spatiotemporal complexity has been developed to understand the ENSO dynamics across a wide range of spatiotemporal scales. The simplest models are the conceptual or box models [5, 6, 7, 8, 9, 10], which involve a few (stochastic) ordinary differential equations. Some variables in these models are directly related to the commonly used low-dimensional ENSO indicators, such as the Niño 3 index. The models at the next level of complexity are called simple dynamical models [11, 12, 13, 14], which consist of a set of partial differential equations to characterize the large-scale dynamics of ENSO. By including more detailed dynamics and coupled relationships with atmosphere and other state variables, the resulting models are named the intermediate complex models (ICMs) [15, 16, 17, 18, 19]. The ICMs are widely used for studying the mechanisms and the forecast of the ENSO. The ENSO dynamics have also been incorporated into the general circulation models (GCMs) [20, 21, 22, 23] to understand its global and regional impacts.

Among these categories of ENSO models, the low-order conceptual models are of particular interest for the following reasons. First, these models are mathematically tractable due to low dimensionality and simple structures. Rigorous analysis of model properties, such as the bifurcations and the stability, can be easily carried out for these models that help understand the mechanisms of the large-scale ENSO features. Second, these models are often utilized to test various hypotheses of physical and causal dependence between different variables. The results can provide valuable guidelines for developing more sophisticated models with refined dynamical structures [7, 24]. These conceptual models have also been used as testbeds for guiding simple dynamical models or ICMs to include additional stochastic noise or stochastic parameterizations [9, 19]. Third, because of the low computational cost, these conceptual models can also be used to predict the ENSO indices [25, 26] and study the predictability from a statistical point of view [27] in light of ensemble forecast methods. Several low-order conceptual models have been independently developed, including the recharge-discharge oscillator [7, 28], the delayed oscillator [29, 16, 30], the western-Pacific oscillator [31], and the advective–reflective oscillator [32]. Later, a unified ENSO oscillator motivated by the dynamics and thermodynamics of Zebiak and Cane’s coupled ocean–atmosphere model has also been built [33]. These models were mainly proposed based on physical intuitions and highlighted one or two specific dynamical features of the ENSO as the building blocks. They have led to many successes in applications.

In this paper, a framework of low-order conceptual models for the ENSO is systematically derived from a spatially-extended stochastic dynamical system with full mathematical rigor. The spatially-extended stochastic dynamical system has a linear, deterministic, and stable dynamical core, describing the interactions between atmosphere, ocean and SST at the interannual time scale. It also exploits a simple stochastic process with state dependent (i.e., multiplicative) noise to parameterize the intraseasonal wind burst activities, including the effect from the Madden-Julian oscillation (MJO), that trigger or terminate the El Niño events. The model captures many observed dynamical features of the ENSO, such as the varying amplitudes and durations of different El Niños as well as the extreme events. It also reproduces the observed power spectrum and non-Gaussian statistics of the SST in the eastern Pacific, the region of the active ENSO events, thanks to the multiplicative noise. Since the dynamical core is linear, a principal component analysis based on the eigenvalue decomposition method is applied to provide a low-order representation of the large-scale dynamics of the ENSO. By further projecting the stochastic wind bursts to the leading order basis functions, the low-order representation succeeds in reproducing the observed large-scale dynamical and statistical features of the ENSO. In light of such a strategy of reducing the model complexity, a low-order conceptual modeling framework is thus rigorously derived. The stochastic versions of many state-of-the-art low-order oscillators, such as the recharge-discharge and the delayed oscillators, become special cases within this framework.

The development of the low-order conceptual modeling framework here is very different from the unified ENSO oscillator [33] and many other models. First, the unified oscillator exploits Zebiak and Cane’s model, which utilizes nonlinearity to trigger the ENSO cycles, as a starting reference model to provide guidelines for developing the conceptual model. In contrast, the conceptual model developed here starts from a linear spatially-extended dynamical system, which uses multiplicative noise to induce the ENSO events. Second, instead of providing only a heuristic building block, the linear dynamical core allows rigorous mathematical derivations from the spatially-extended dynamical system to the low-order conceptual model. This procedure facilitates rigorous analysis and direct intercomparison between the two systems. Third, the low-order conceptual modeling framework here provides outputs for all the atmosphere, ocean, and SST components with detailed spatiotemporal patterns. This contrasts with many existing conceptual models, which focus only on a small set of specified state variables and often involve the averaged quantities over a large domain as a coarse-graining representation. Despite the intrinsic low dimensionality of the model, the state variables for atmosphere, ocean, and SST in this conceptual modeling framework are connected via the eigenvectors. Therefore, depending on the quantity of interest, the model can explicitly characterize the relationships between different atmosphere-ocean-SST components contributing to the ENSO dynamics.

The rest of the paper is organized as follows. Section 2 describes the observational data. Section 3 presents the simple spatially-extended stochastic dynamical model and its properties. Section 4 utilizes the eigenvalue decomposition to illustrate the reduced order representation of the ENSO dynamics. Section 5 shows the mathematical expressions of the conceptual models, including the recharge-discharge and the delayed oscillators. The paper is concluded in Section 6.

2 Observational Data

The following observational data are utilized in this study. Daily sea surface temperature (SST) data comes from the OISST reanalysis [34] (https://www.ncdc.noaa.gov/oisst). Monthly thermocline depth is from the NCEP GODAS reanalysis [35] (http://www.esrl.noaa.gov/psd/). Note that thermocline depth is computed from potential temperature as the depth of the 20o20^{o}C isotherm. Daily zonal winds at 850 hPa are from the NCEP–NCAR reanalysis [36] (http://www.esrl.noaa.gov/psd/). All datasets are averaged meridionally within 5o5^{o}S-5o5^{o}N in the tropical Pacific (120o120^{o}E–80o80^{o}W). Only the anomalies are presented in this study, calculated by removing the monthly mean climatology of the whole period. The eastern and western Pacific regions are defined as 120o120^{o}E-160o160^{o}W and 160o160^{o}W-80o80^{o}W, respectively. The Niño 3 SST index is the average of SST anomalies over the region of 150o150^{o}W-90o90^{o}W. Denote by TET_{E} the time series of SST averaged over the eastern Pacific. The time series TET_{E} and Niño 3 SST index are almost the same as each other.

3 The Simple Spatially-Extended Stochastic Model

3.1 Review of the model

The simple spatially-extended stochastic model was originally developed in [13]. The starting system is a coupled atmosphere-ocean-SST model, which is given by

Atmosphere:

yvxθ=0yuyθ=0(xu+yv)=Eq/(1Q¯)\begin{split}&-yv-\partial_{x}\theta=0\\ &yu-\partial_{y}\theta=0\\ &-(\partial_{x}u+\partial_{y}v)=E_{q}/(1-\overline{Q})\end{split} (1)

Ocean:

tUc1YV+c1xH=c1τYU+YH=0tH+c1(xU+YV)=0\begin{split}&\partial_{t}U-c_{1}YV+c_{1}\partial_{x}H=c_{1}{\tau}\\ &YU+\partial_{Y}H=0\\ &\partial_{t}H+c_{1}(\partial_{x}U+\partial_{Y}V)=0\end{split} (2)

SST:

tT=c1ζEq+c1ηH.\begin{split}&\partial_{t}T=-c_{1}\zeta E_{q}+c_{1}\eta H.\end{split} (3)

The above coupled system consists of a non-dissipative Matsuno–Gill type atmosphere model [37, 38], a simple shallow-water ocean model [39] and an SST budget equation [24]. The non-dissipative atmosphere is also consistent with the skeleton model for the MJO in the tropics [40]. In (1)–(3), uu and vv are the zonal and meridional wind speeds, θ\theta is the potential temperature, UU and VV are the zonal and meridional ocean currents, HH is the thermocline depth, TT is the SST, Eq=αqTE_{q}=\alpha_{q}T is the latent heat, and τ=γu\tau=\gamma u is the wind stress along the zonal direction. All the variables are the anomalies from the climatology mean states. There parameters c1c_{1}, Q¯\overline{Q} and ζ\zeta are all constants. The parameter c1c_{1} is a non-dimensional number, representing the ratio of ocean and atmosphere phase speeds over the Froude number, Q¯\overline{Q} is related to the background moisture, and ζ\zeta is the latent heating exchange coefficient. The thermocline feedback η(x)\eta(x) is a fixed spatial dependent function stronger in the eastern Pacific due to the shallower thermocline. See Panel (c) of Figure 1. Also, see the Appendix for a summary of model parameters. Note that different axes in the meridional directions are utilized for the atmosphere (yy) and ocean (YY). This is because the deformation radii of the atmosphere and the ocean are different. The atmosphere covers the entire equatorial band from [0,LA][0,L_{A}], and therefore the periodic boundary condition in the zonal direction is applied, namely u(0,y,t)=u(LA,y,t)u(0,y,t)=u(L_{A},y,t). The ocean covers the equatorial Pacific [0,LO][0,L_{O}] with the boundary conditions U(0,Y,t)dY=0\int_{-\infty}^{\infty}U(0,Y,t){\,\rm d}Y=0 and U(LO,Y,t)=0U(L_{O},Y,t)=0 [41, 24].

Since the most predominant behavior of the ENSO is around the equator and along the zonal direction, it is natural to apply a meridional expansion, which serves as the separation of variables, and then implement a meridional truncation to the first basis in both atmosphere ϕ0(Y)\phi_{0}(Y) and ocean ψ0(y)\psi_{0}(y) to reduce the complexity of the system. The basis functions in the meridional direction are given by the parabolic cylinder functions [42]. See Panels (a)–(b) in Figure 1 for the meridional basis functions, where the first one has a Gaussian profile. The meridional truncation triggers atmosphere Kelvin, Rossby waves KA,RAK_{A},R_{A}, and ocean Kelvin, Rossby waves KO,ROK_{O},R_{O}. Therefore, after removing the meridional dependence from (1)–(3), the coupled system (with dependence only on time tt and zonal coordinate xx) yields.

Atmosphere:

xKA=χAEq(22Q¯)1xRA/3=χAEq(33Q¯)1(B.C.)KA(0,t)=KA(LA,t)(B.C.)RA(0,t)=RA(LA,t)\begin{split}&\partial_{x}K_{A}=-\chi_{A}E_{q}(2-2\bar{Q})^{-1}\\ &-\partial_{x}R_{A}/3=-\chi_{A}E_{q}(3-3\bar{Q})^{-1}\\ (B.C.)~{}~{}~{}&K_{A}(0,t)=K_{A}(L_{A},t)\\ (B.C.)~{}~{}~{}&R_{A}(0,t)=R_{A}(L_{A},t)\end{split} (4)

Ocean:

tKO+c1xKO=χOc1τ/2tRO(c1/3)xRO=χOc1τ/3(B.C.)KO(0,t)=rWRO(0,t)(B.C.)RO(LO,t)=rEKO(LO,t)\begin{split}&\partial_{t}K_{O}+c_{1}\partial_{x}K_{O}=\chi_{O}c_{1}\tau/2\\ &\partial_{t}R_{O}-(c_{1}/3)\partial_{x}R_{O}=-\chi_{O}c_{1}\tau/3\\ (B.C.)~{}~{}~{}&K_{O}(0,t)=r_{W}R_{O}(0,t)\\ (B.C.)~{}~{}~{}&R_{O}(L_{O},t)=r_{E}K_{O}(L_{O},t)\end{split} (5)

SST:

tT=c1ζEq+c1η(KO+RO).\begin{split}&\partial_{t}T=-c_{1}\zeta E_{q}+c_{1}\eta(K_{O}+R_{O}).\end{split} (6)

The boundary conditions (B.C.) in (4)–(6) are derived from those related to (1)–(3). Periodic boundary conditions are remained for atmosphere waves, while those for the ocean velocity become the reflection boundary conditions for ocean waves. The constants χA\chi_{A} and χO\chi_{O} are the meridional projection coefficients with χA=ϕ0(y)ϕ0(y/c)𝑑y\chi_{A}=\int_{-\infty}^{\infty}\phi_{0}(y)\phi_{0}(y/\sqrt{c})dy and χO=ψ0(Y)ψ0(cY)𝑑Y\chi_{O}=\int_{-\infty}^{\infty}\psi_{0}(Y)\psi_{0}(\sqrt{c}Y)dY, where cc is the ratio of ocean and atmosphere phase speed. Once these waves are solved, the physical variables can be reconstructed,

u=(KARA)ϕ0+(RA/2)ϕ2θ=(KA+RA)ϕ0(RA/2)ϕ2U=(KORO)ψ0+(RO/2)ψ2H=(KO+RO)ψ0+(RO/2)ψ2\begin{split}&u=(K_{A}-R_{A})\phi_{0}+(R_{A}/\sqrt{2})\phi_{2}\\ &\theta=-(K_{A}+R_{A})\phi_{0}-(R_{A}/\sqrt{2})\phi_{2}\\ &U=(K_{O}-R_{O})\psi_{0}+(R_{O}/\sqrt{2})\psi_{2}\\ &H=(K_{O}+R_{O})\psi_{0}+(R_{O}/\sqrt{2})\psi_{2}\end{split} (7)

where ϕ2\phi_{2} and ψ2\psi_{2} are the third meridional bases of atmosphere and ocean, respectively, resulting from the meridional expansion of the parabolic cylinder function [42]. See Panels (a)–(b) in Figure 1 for the structures of these bases.

In addition to the interannual state variables, the intraseasonal variabilities are also part of the indispensable ENSO dynamics. Although these intraseasonal variabilities behave like random noise in the interannual time scale, they play vital roles in generating the ENSO events and increasing the ENSO complexity. In particular, atmosphere wind bursts in the intraseasonal time scale, such as the westerly wind bursts (WWBs) [43, 44, 45], easterly wind bursts (EWBs) [46, 47] and the MJO [48, 49], are essential in triggering and terminating the El Niño events. Due to the fast dynamics of wind activities, it is natural to adopt a simple stochastic process to effectively characterize their time evolutions. With the stochastic wind bursts, the zonal wind stress τ\tau now has two components, τ=γ(u+up)\tau=\gamma(u+u_{p}), where uu is directly from the atmosphere model (4) while upu_{p} is the contribution from the stochastic wind bursts. Here,

up=ap(t)sp(x)ϕ0(y),u_{p}=a_{p}(t)s_{p}(x)\phi_{0}(y), (8)

where for simplicity sp(x)s_{p}(x) is a fixed spatial basis function being localized in the western Pacific (see Panel (c) of Figure 1), since most of the active wind bursts are observed there. The stochastic amplitude apa_{p} is given by the following stochastic differential equation

dapdt=dpap+σp(TW)W˙(t),\frac{{\,\rm d}a_{p}}{{\,\rm d}t}=-d_{p}a_{p}+{\sigma_{p}(T_{W})}\dot{W}(t), (9)

where W˙(t)\dot{W}(t) is a standard white noise [50] and dpd_{p} is a damping term characterizing the temporal correlation of the wind bursts at the intraseasonal time scale.

Note that the noise coefficient σp(TW){\sigma_{p}(T_{W})} is state-dependent (namely, the multiplicative noise) as a function of the averaged SST in the western Pacific. This is because a warmer SST induces more convective activities, which in turn triggers more wind bursts [44, 43]. Here, the state dependence of σp(TW){\sigma_{p}(T_{W})} is given by a two-state Markov jump process with one active state and one quiescent state:

σp(TW)={σp0=0.2 for quiescent state 0σp1=2.6 for active state 1\sigma_{p}\left(T_{W}\right)=\left\{\begin{array}[]{l}\sigma_{p0}=0.2\quad\text{ for quiescent state }0\\ \sigma_{p1}=2.6\quad\text{ for active state }1\end{array}\right. (10)

The transition rate from the quiescent to the active state μ01\mu_{01} and that from the active to the quiescent state μ10\mu_{10} are shown in Panel (d) of Figure 1. The explicit equations of them are in the following:

State 1 to 0:μ10=14(1tanh(2TW))State 0 to 1:μ01=18(tanh(2TW)+1)\begin{split}&\text{State 1 to 0:}\ \mu_{10}=\frac{1}{4}\left(1-\tanh\left(2T_{W}\right)\right)\\ &\text{State 0 to 1:}\ \mu_{01}=\frac{1}{8}\left(\tanh\left(2T_{W}\right)+1\right)\end{split} (11)

Here, transition rates μ01\mu_{01} and μ10\mu_{10} are correlated and anti-correlated with TWT_{W}, respectively, which reflect the nature of the multiplicative noise. Note that the wind burst activity from the model does not favor specifically westerly or easterly. The WWBs and EWBs generated from the model are entirely random. Only the amplitude of the wind bursts is determined by the transitions related to the SST. The two-state Markov jump process for parameterizing the wind activities is utilized here for simplicity, categorizing the winds by only the active and quiescent phases. A continuous dependence of the wind strength on the SST can easily be incorporated, which leads to a similar SST behavior in the ENSO dynamics [51, 52].

As a final remark, the coupled model (4)–(9) focuses on the eastern Pacific El Niño events, which is the primary interest of this study. The model does not include the mechanism of generating the central Pacific events [53, 54, 55]. Incorporating the central Pacific events into the model will be briefly discussed in Section 6.

3.2 Model properties

The dynamical core of the system (4)–(6) is deterministic, linear and stable. In other words, the solution is a linear damped regular oscillator. The random wind bursts serve as external forcing to trigger the El Niño events and introduce irregularity in the solution. The subsequent La Niña is a consequence of the decaying phase of the ENSO cycle.

Panel (e) in Figure 1 shows one simulation of the coupled model (4)–(9) for different fields. The model can reproduce the quasi-regular oscillation behavior of the SST anomaly. When the SST in the western Pacific increases, the wind activities become more significant. If the strong wind activity is dominated by the WWBs, then El Niño events are likely to be triggered. During this process, the center of the atmosphere wind convergence moves towards the east, and the strong eastward ocean zonal current is observed in the western Pacific. The thermocline depth becomes deeper in the eastern Pacific and induces the El Niño events. The model can generate not only the moderate El Niño events, but also the extreme events, known as the super El Niños, that have strong amplitudes (e.g., the event at t=490t=490). In addition to reproducing the super El Niño that mimics the observed 1997-1998 event (e.g., at t=490t=490), the model succeeds in generating the so-called delayed super El Niño [56], similar to the observed 2014-2016 event [57, 47]. One such example is the event around t=474t=474-476476, which is due to the peculiar wind structures. A series of WWBs induce the onset of an El Niño event. However, a strong EWB occurs immediately to present the development of such an event to a super El Niño. After a year, another period of strong WWBs happen to trigger the extreme event.

Averaging the SST over the eastern Pacific, the resulting time series is denoted by TET_{E}. The model can reproduce very similar non-Gaussian probability density function (PDF) of TET_{E} as the observations [13]. This non-Gaussian feature is reproduced with the contribution from the multiplicative noise, without which the model is linear and the associated PDF is Gaussian. Likewise, the spectrum of the model has the most significant power within the band of 3 to 7 years, which is also consistent with observations.

Refer to caption
Figure 1: Structure and simulation of the spatially-extended stochastic system. Panels (a)–(b): meridional basis functions of the atmosphere (ϕ0,ϕ2\phi_{0},\phi_{2}) and ocean (ψ0,ψ2\psi_{0},\psi_{2}). Panel (c): Zonal profile of the thermocline depth η(x)\eta(x) and wind burst region sp(x)s_{p}(x). Panel (d): the transition rates in the two-state Markov jump process. Panel (e): a random realization of the model simulation. The variables are the atmosphere zonal wind velocity uu (unit: m/s), the ocean zonal current UU (unit: m/s), the thermocline depth HH (unit: m), SST TT (unit: oC), ocean Kelvin wave KOK_{O} (no unit), ocean Rossby wave ROR_{O} (no unit), the state of the Markov jump process, the wind burst time series ap(t)a_{p}(t) (unit: m/s), and the two SST indices TWT_{W} and TET_{E} (unit: oC), which are the averaged value over the western and eastern Pacific, respectively. In the panel of the wind burst amplitude ap(t)a_{p}(t), a 90-day running average of ap(t)a_{p}(t) (brown curve) is included to represent the lower frequency part of the wind bursts.

4 Reduced-Order Representation of the ENSO dynamics

4.1 Discretization of the deterministic, linear and stable dynamics

Without the stochastic wind bursts, the coupled atmosphere-ocean-SST model (4)–(6) is a linear, deterministic and stable model. Since uu (and therefore the wind stress τ\tau) is a function of KAK_{A} and RAR_{A}, the coupled system can be rewritten as

tKO+c1xKO\displaystyle\partial_{t}{K_{O}}+c_{1}\partial_{x}{K_{O}} =a2(KARA),\displaystyle=\frac{a}{2}\ ({K_{A}}-{R_{A}}), KO(0,t)=rWRO(0,t),\displaystyle K_{O}(0,t)=r_{W}R_{O}(0,t), (12)
tROc13xRO\displaystyle\partial_{t}{R_{O}}-\frac{c_{1}}{3}\partial_{x}{R_{O}} =a3(KARA),\displaystyle=-\frac{a}{3}\ ({K_{A}}-{R_{A}}), RO(LO,t)=rEKO(LO,t),\displaystyle R_{O}(L_{O},t)=r_{E}K_{O}(L_{O},t),
tT\displaystyle\partial_{t}{T} =bT+c1η(KO+RO),\displaystyle=-b{T}+c_{1}\eta({K_{O}}+{R_{O}}),
dAKA+xKA\displaystyle d_{A}{K_{A}}+\partial_{x}{K_{A}} =m1αq(𝟏[0,LO]αqTEq),\displaystyle=\frac{m_{1}}{\alpha_{q}}(\mathbf{1}_{[0,L_{O}]}\alpha_{q}{T}-\langle{E_{q}}\rangle), KA(0,t)=KA(LA,t),\displaystyle K_{A}(0,t)=K_{A}(L_{A},t),
dARA13xRA\displaystyle d_{A}{R_{A}}-\frac{1}{3}\partial_{x}{R_{A}} =m2αq(𝟏[0,LO]αqTEq),\displaystyle=\frac{m_{2}}{\alpha_{q}}(\mathbf{1}_{[0,L_{O}]}\alpha_{q}{T}-\langle{E_{q}}\rangle), RA(0,t)=RA(LA,t),\displaystyle R_{A}(0,t)=R_{A}(L_{A},t),

where a small damping dA=108d_{A}=10^{-8} is added to guarantee the solution is unique [13], and the new constants introduced in (12) are the following ones:

a=χOc1γ,b=c1ζαq,m1=χAαq/(22Q¯),m2=χAαq/(33Q¯).a=\chi_{O}c_{1}\gamma,\quad b=c_{1}\zeta\alpha_{q},\quad m_{1}=-\chi_{A}\alpha_{q}/(2-2\bar{Q}),\quad m_{2}=-\chi_{A}\alpha_{q}/(3-3\bar{Q}).

The domain of KO,ROK_{O},R_{O} and TT is the equatorial Pacific ocean [0,LO][0,L_{O}] while the domain of KAK_{A} and RAR_{A} is the entire equatorial band [0,LA][0,L_{A}].

Define a vector 𝐮\mathbf{u} that contains all the prognostic variables at discrete equal-partitioned grid points,

𝐮=(𝐊𝐎;𝐑𝐎;𝐓)=(KO,1,KO,NO,RO,1,RO,NO,T1,TNO)𝚃,\mathbf{u}=(\mathbf{K_{O}};\mathbf{R_{O}};\mathbf{T})=(K_{O,1},\ldots K_{O,{N_{O}}},R_{O,1},\ldots R_{O,{N_{O}}},T_{1},\ldots T_{N_{O}})^{\mathtt{T}}, (13)

where NON_{O} is the total number of grid points over the equatorial Pacific region [0,LO][0,L_{O}] and 𝚃\cdot^{\mathtt{T}} is the vector transpose. Using an upwind scheme, the discrete form of the linear system (12) is given by:

d𝐮dt=𝐌𝐮.\frac{{\,\rm d}\mathbf{u}}{{\,\rm d}t}=\mathbf{Mu}. (14)

The detailed form of 𝐌\mathbf{M} is included in the Appendix. With the stochastic wind burst, the above equation can be formally written as:

d𝐮dt=𝐌𝐮+𝐅,\frac{{\,\rm d}\mathbf{u}}{{\,\rm d}t}=\mathbf{Mu}+\mathbf{F}, (15)

where 𝐅\mathbf{F} is associated with the stochastic forcing. Denote by 𝐬𝐮=(χOc1γsp/2,χOc1γsp/3,𝟎)𝚃\mathbf{s}_{\mathbf{u}}=(\chi_{O}c_{1}\gamma s_{p}/2,\\ -\chi_{O}c_{1}\gamma s_{p}/3,\mathbf{0})^{\mathtt{T}} and then 𝐅=𝐬𝐮ap\mathbf{F}=\mathbf{s}_{\mathbf{u}}a_{p}, where sps_{p} is written in the discrete form at grid points 1,,NO1,\ldots,N_{O} and 𝟎\mathbf{0} is a 1×NO1\times N_{O} vector with all zero entries. This is the wind burst forcing acting on the three components 𝐮=(𝐊O,𝐑O,𝐓)\mathbf{u}=(\mathbf{K}_{O},\mathbf{R}_{O},\mathbf{T}) (ocean Kelvin wave, ocean Rossby wave, and SST) of the state variables. See Panel (a) of Figure 2 for the profile of 𝐬𝐮\mathbf{s}_{\mathbf{u}}. As expected, a positive response is on the ocean Kelvin waves and a negative one is on the ocean Rossby waves, both in the western Pacific. The forcing is not directly imposed on SST; therefore, the third segment of 𝐬𝐮\mathbf{s}_{\mathbf{u}}, corresponding to SST components, is zero.

Refer to caption
Figure 2: Zonal structure of the functions 𝐬𝐮\mathbf{s}_{\mathbf{u}} and 𝐳\mathbf{z}. The three segments correspond to the three components 𝐮=(𝐊O,𝐑O,𝐓)\mathbf{u}=(\mathbf{K}_{O},\mathbf{R}_{O},\mathbf{T}) (ocean Kelvin wave, ocean Rossby wave, and SST).

4.2 The eigenvalues of the system

It can be numerically validated that the 3NO3N_{O} eigenvalues all have distinct values and are non-zero. In addition, the ii-th left eigenvector and the jj-th right eigenvector of the matrix 𝐌\mathbf{M} are orthogonal with each other for iji\neq j. Therefore, in light of the matrix similarity, the following result is arrived at:

𝐗1𝐌𝐗=𝚲,\mathbf{X}^{-1}\mathbf{M}\mathbf{X}=\mathbf{\Lambda}, (16)

or equivalently

𝐌=𝐗𝚲𝐗1,\mathbf{M}=\mathbf{X}\mathbf{\Lambda}\mathbf{X}^{-1}, (17)

where 𝚲\mathbf{\Lambda} is a diagonal matrix with diagonal entries being eigenvalues and 𝐗\mathbf{X} contains the 3NO3N_{O} right eigenvectors. Since the right eigenvectors 𝐗\mathbf{X} are complete, the matrix inverse exists.

Panel (a) of Figure 3 shows the eigenvalues associated with the 3NO×3NO3N_{O}\times 3N_{O} matrix 𝐌\mathbf{M}, where NO=56N_{O}=56 is utilized here and in the subsequent numerical simulations. These eigenvalues are complex numbers. The real and imaginary parts are shown in the y- and x-axis, representing the growth rate and the frequency, respectively. Since the starting system is stable, the growth rates of all the eigenvalues are negative, representing the decay of the solution. These eigenvalues are ranked according to the growth rate. Therefore, the leading two eigenmodes (marked by red dots) are the dominant modes of the system since they have the slowest decaying rates. These two modes appear in pairs; both the eigenvalues and the eigenvectors are complex conjugates. Their eigenvalues have a frequency of 4.64.6 years and a decay rate of 1.51.5 years, which are consistent with the observed features of ENSO. The reconstructed spatiotemporal patterns associated with these two eigenmodes (by setting the growth rate to be zero for the plotting purpose), as shown in Panel (b) of (3), mimic the large-scale ENSO oscillation solution. Notably, the leading two eigenmodes are very robust with respect to the spatial resolution NON_{O} as long as NON_{O} is not too small. Therefore, the analysis here provides a theoretical justification for developing a low-order model based on these leading two eigenmodes.

Refer to caption
Figure 3: Eigenvalue analysis of the spatially-extended stochastic system. Panel (a): the first a few eigenvalues. The eigenvalues are complex numbers. The real and imaginary parts are shown in the y- and x-axis, representing the growth rate and the frequency, respectively. These eigenvalues are ranked according to the growth rate. The first two eigenvalues, which are a pair of complex conjugates, are marked in red color. Panel (b): The reconstructed linear oscillation solution from the leading two eigenmodes, which is given by taking the summation of the two products of each of the eigenvalues and the associated eigenvectors. The growth rate is set to be zero here to display the oscillation patterns.

4.3 Comparison of the projected solution with the full solution

Before developing a reduced-order conceptual model based on the eigenvalue analysis, as shown above, it is essential to study the similarity and the gap between the full solution from the spatially-extended stochastic system (4)–(9) and that from the low-dimensional representation consisting of only the leading two eigenmodes, namely the projected solution.

Figure 4 includes a comparison of the SST between these two solutions. The projected solution (Panel (b)) is qualitatively similar to the full solution (Panel (a)) with an almost identical pattern and a nearly negligible residual error (Panel (c)). This is not surprising since the decaying rates of the remaining eigenmodes are more significant than those of the leading two components. Thus, the time evolution of the signal is overall explained by the two leading modes. In addition to the spatiotemporal pattern, the two key statistics, namely the power spectrum and the PDF, of the averaged SST over the eastern Pacific TET_{E} are almost indistinguishable between the full and the projected solutions. In particular, the power spectrum peaks between 3 and 7 years, which is consistent with the observed frequency of the ENSO cycles. The projected solution also recovers the non-Gaussian PDF of TET_{E}, which has almost the same skewness and kurtosis as the full solution. Non-Gaussianity is a crucial feature of the ENSO statistics, which indicates a non-symmetry between the El Niño and the La Niña. The fat tail on the positive side corresponds to the super El Niños, which remains in the projected solution.

Refer to caption
Figure 4: Comparison of the full solution from the spatially-extended stochastic system (4)–(9) with that from the low-order representation consisting of only the leading two eigenmodes, namely the projected solution. Panels (a)–(c): One realization of the full solution and the corresponding projected solution as well as their difference, namely the residual. Panels (d)–(g): the statistics of the time series TET_{E}, which is the averaged SST over the eastern Pacific. Panels (d)–(e) compare the spectrum, where the two regions bounded by the two dashed black curves represent the 3- to the 7-year window. A running average showing the smoothed version of the spectrum is also included. Panels (f)–(g) compare the PDFs, where the red dashed curve shows the Gaussian fit. The values of the skewness and the kurtosis of TET_{E} are also included to indicate the strong non-Gaussian features.

Figure 5 includes a more detailed comparison between the full and the projected solutions by showing different physical and wave variables. A delayed super El Niño event (from t=474t=474 to t=476t=476) occurs during this period. Overall, the projected solution captures the spatiotemporal patterns of the full one. Especially, the delayed super El Niño is reproduced in the projected solution. Yet, the small-scale wave propagation is missed in the projected solution. This is particularly unambiguous in the panels of the ocean Kelvin wave KOK_{O} and the ocean Rossby wave ROR_{O}. As a result, the values across the western Pacific or the eastern Pacific remain constant for different fields. This is perhaps the most significant difference between the full and the projected solutions. Small-scale wave propagations are expected to appear at least with three degrees of freedom (e.g., either three variables or three regions across Pacific) [9]. Nevertheless, the projected solution clearly captures the large-scale eastward moving patterns, representing the ENSO development.

Refer to caption
Figure 5: A case study of the delayed super El Niño event (from t=474t=474 to t=476t=476) for comparing the full solution from the spatially-extended stochastic system with that from the low-order representation consisting of only the leading two eigenmodes, namely the projected solution. The variables are the atmosphere zonal wind velocity uu (unit: m/s), the ocean zonal current UU (unit: m/s), the thermocline depth HH (unit: m), SST TT (unit: oC), ocean Kelvin wave KOK_{O} (no unit), ocean Rossby wave ROR_{O} (no unit), the state of the Markov jump process, and the wind burst time series ap(t)a_{p}(t) (unit: m/s). In the panel of the wind burst amplitude ap(t)a_{p}(t), a 90-day running average of ap(t)a_{p}(t) (brown curve) is included to represent the lower frequency part of the wind bursts.

Figure 6 compares the full observed SST using the reanalysis data (Panel (a)) with the projected data using the leading two eigenmodes computed from the model. Panel (b) shows the results by directly projecting the SST to these two modes. The projected solution resembles the full observations in the eastern Pacific with a pattern correlation above 0.850.85 at all longitude grid points. In contrast, the projected solution in the central-western Pacific differs from the full solution. This is not surprising since the starting model and the low-order presentation here focus only on the eastern Pacific El Niño events and do not have the mechanism to create the central Pacific events. The projected solution in the western Pacific captures the overall patterns of the full observations, though the details are somewhat missed. In addition to using the SST data, the atmosphere wind data can also be used to reconstruct the low-order representation of the SST. See Panel (c). Here, the atmosphere wind data is first projected to the leading two eigenvectors corresponding to the atmosphere components to compute the projection coefficients. Then these coefficients are multiplied by the eigenvectors corresponding to the SST components for reconstruction. The reconstructed low-order representation of SST using the atmosphere reanalysis data has a similar pattern as that based on the direct projection using the SST in the eastern Pacific. Finally, Panel (d) shows the reconstructed SST data using a direct data-driven regression method. This is achieved by first computing the correlation coefficient r(x)r(x) between the SST at each longitude with the Niño 3 SST index using the observational data and then carrying out the spatiotemporal reconstruction using the formula SST(x,t)=r(x)Niño 3(t)\mbox{SST}(x,t)=r(x)\mbox{Ni\~{n}o 3}(t). Note that this procedure uses the entire observational data and can be regarded as one of the optimal linear reconstruction methods. It serves as a benchmark solution. The results show that, similar to the low-dimensional projected solutions in Panels (b)–(c), the reconstruction in the eastern Pacific resembles the truth while that in the central-western Pacific contains a significant error. The comparison between the results in Panels (b)–(c) and Panel (d) implies that the reconstruction using the model-induced basis functions is skillful in reproducing the eastern Pacific El Niño events. It also indicates that both the wind and the SST data can be used for the reconstruction, allowing large freedom in choosing the available data sets.

Refer to caption
Figure 6: Comparison of the observational SST data (Panel (a)) with low-order representations of the spatiotemporal solution using different reconstructed methods. Panel (b): the reconstruction by projecting the observed SST to the leading two eigenmodes from the spatially-extended system. Panel (c): the reconstruction by projecting the observed wind data to the leading two eigenmodes from the spatially-extended system and then recovering the SST patterns via the interconnection between the wind and the SST components in the eigenvectors. Panel (d): the reconstruction by using a linear regression to determine the SST. The regression coefficient is determined by computing the correlation coefficient between the SST at each longitude with the Niño 3 SST index using the observational data. In Panels (b)–(d), the residual in the reconstruction, namely the difference between the truth shown in Panel (a) and the reconstructed spatiotemporal SST field, is also shown. In addition, the correlation between the reconstructed SST field with the truth at each longitude is included beneath the reconstructed field.

5 Reduced-Order Conceptual Models

Given the fact that the reconstructed SST fields from the projected solution and the full solution are similar to each other, it is natural to develop a reduced-order conceptual model using these two eigenmodes to characterize ENSO complexity with low computational cost.

5.1 The general framework

Plugging the relationship of the matrix decomposition 𝐌=𝐗𝚲𝐗1\mathbf{M}=\mathbf{X}\mathbf{\Lambda}\mathbf{X}^{-1} in (17) into the full system (15) yields

d𝐮dt=𝐗𝚲𝐗1𝐮+𝐅,\frac{{\,\rm d}\mathbf{u}}{{\,\rm d}t}=\mathbf{X}\mathbf{\Lambda}\mathbf{X}^{-1}\mathbf{u}+\mathbf{F}, (18)

where 𝚲\mathbf{\Lambda} is a diagonal matrix. Then multiplying (18) by 𝐗1\mathbf{X}^{-1} gives 3NO3N_{O} independent equations

d𝐯dt=𝚲𝐯+𝐆,\frac{{\,\rm d}\mathbf{v}}{{\,\rm d}t}=\mathbf{\Lambda}\mathbf{v}+\mathbf{G}, (19)

where

𝐯=𝐗1𝐮,and𝐆=𝐗1𝐅.\mathbf{v}=\mathbf{X}^{-1}\mathbf{u},\qquad\mbox{and}\qquad\mathbf{G}=\mathbf{X}^{-1}\mathbf{F}. (20)

Once 𝐯\mathbf{v} is solved, 𝐮\mathbf{u} can be obtained by

𝐮=𝐗𝐯.\mathbf{u}=\mathbf{X}\mathbf{v}. (21)

Denote by 𝐳𝚃\mathbf{z}^{\mathtt{T}} and 𝐳¯𝚃\bar{\mathbf{z}}^{\mathtt{T}} the first two rows of 𝐗1\mathbf{X}^{-1}. See Panel (b) of Figure 2 for the profile of 𝐳\mathbf{z}. Likewise, denote by 𝐱\mathbf{x} and 𝐱¯\bar{\mathbf{x}} the first two columns of 𝐗\mathbf{X}. The corresponding leading two eigenvalues are denoted by λo\lambda_{o} and λ¯o\bar{\lambda}_{o} (the two red dots in Panel (a) of Figure 3), which are written as do±iωo-d_{o}\pm i\omega_{o} with do<0-d_{o}<0 and ωo\omega_{o} being the growth rate and the frequency, respectively. Here 𝚃\cdot^{\mathtt{T}} is the transpose, and ¯\bar{\cdot} is the complex conjugate. Clearly, 𝐳𝚃𝐱=𝐳¯𝚃𝐱¯=1\mathbf{z}^{\mathtt{T}}\mathbf{x}=\bar{\mathbf{z}}^{\mathtt{T}}\bar{\mathbf{x}}=1.

Proposition 5.1.

Define v=𝐳𝚃𝐮v=\mathbf{z}^{\mathtt{T}}\mathbf{u} and g=𝐳𝚃𝐅=𝐳𝚃𝐬𝐮apg=\mathbf{z}^{\mathtt{T}}\mathbf{F}=\mathbf{z}^{\mathtt{T}}\mathbf{s}_{\mathbf{u}}a_{p} and naturally, v¯=𝐳¯𝚃𝐮\bar{v}=\bar{\mathbf{z}}^{\mathtt{T}}\mathbf{u} and g¯=𝐳¯𝚃𝐅=𝐳¯𝚃𝐬𝐮ap\bar{g}=\bar{\mathbf{z}}^{\mathtt{T}}\mathbf{F}=\bar{\mathbf{z}}^{\mathtt{T}}\mathbf{s}_{\mathbf{u}}a_{p}. In light of the first two dimensions of (19), the following two-dimensional forced oscillator system is reached,

dvdt=λov+g,dv¯dt=λ¯ov¯+g¯.\begin{split}\frac{{\,\rm d}v}{{\,\rm d}t}&=\lambda_{o}v+g,\\ \frac{{\,\rm d}\bar{v}}{{\,\rm d}t}&=\bar{\lambda}_{o}\bar{v}+\bar{g}.\end{split} (22)

Once vv and v¯\bar{v} are computed, the approximate of 𝐮\mathbf{u} based on these two modes (namely the low-order representation of the spatiotemporal reconstructed solution), which is denoted by 𝐮~\widetilde{\mathbf{u}}, can be computed via

𝐮~=v𝐱+v¯𝐱¯.\widetilde{\mathbf{u}}=v\mathbf{x}+\bar{v}\bar{\mathbf{x}}. (23)

Figure 7 shows the eigenvector associated with λo\lambda_{o} for different model components.

It is worthwhile to highlight that, as was shown in (13), the solution 𝐮~\widetilde{\mathbf{u}} consists of the ocean Kelvin wave KOK_{O}, the ocean Rossby wave ROR_{O}, and the SST TT. Recall in (7) that the ocean current UU and the thermocline depth HH are the linear combinations of KOK_{O} and ROR_{O}. In addition, the atmosphere wind uu and the potential temperature θ\theta can be solved by exploiting the linear relationship with the ocean and SST variables according to (12) (see Appendix for details). Therefore, the two-dimensional low-order conceptual model (22) provides the solutions for all the atmosphere, ocean, and SST fields with detailed spatiotemporal solutions. This contrasts with many existing conceptual models, which focus only on a small set of specified state variables and often merely involve the averaged quantities over a large domain.

Refer to caption
Figure 7: The eigenvector associated with λo\lambda_{o} corresponds to different variable components.

5.2 Two-dimensional forced oscillator with arbitrarily specified quantities as state variables

The general framework in (22) also allows the development of the reduced-order models with arbitrarily specified quantities (in addition to the direct model output uu, θ\theta, HH, UU and TT) as state variables. This can be achieved by introducing two new real-valued state variables a~\widetilde{a} and b~\widetilde{b}, which are the linear combinations of vv and v¯\bar{v} by angle rotations and amplifications,

a~\displaystyle\widetilde{a} =(eiαv+eiαv¯)Aa,\displaystyle=\left(e^{i\alpha}v+e^{-i\alpha}\bar{v}\right)A_{a}, (24a)
b~\displaystyle\widetilde{b} =(eiβv+eiβv¯)Ab.\displaystyle=\left(e^{i\beta}v+e^{-i\beta}\bar{v}\right)A_{b}. (24b)

Here α\alpha and β\beta are rotating angles to be in phase with the two new physical variables of interest while the real variables AaA_{a} and AbA_{b} are the amplitudes. Here, it is required that α,β[π,π]\alpha,\beta\in[-\pi,\pi] and βα\beta\neq\alpha to rule out the degenerated case. The two variables a~\widetilde{a} and b~\widetilde{b} can be defined as any variables that can be represented by a linear combination of the model state variables.

Proposition 5.2.

The approximate solution of 𝐮\mathbf{u} based on the two arbitrary state variables a~\widetilde{a} and b~\widetilde{b} is given by

𝐮~=𝐱a~a~+𝐱b~b~,{\widetilde{\mathbf{u}}}=\mathbf{x}_{\tilde{a}}\widetilde{a}+\mathbf{x}_{\tilde{b}}\widetilde{b}, (25)

where

𝐱a~=1Aa(eiα𝐱ei2αei2β+eiα𝐱¯ei2αei2β),𝐱b~=1Ab(eiβ𝐱ei2αei2β+eiβ𝐱¯ei2αei2β)\begin{split}\mathbf{x}_{\tilde{a}}&=\frac{1}{A_{a}}\left(\frac{e^{i\alpha}\mathbf{x}}{e^{i2\alpha}-e^{i2\beta}}+\frac{e^{-i\alpha}\bar{\mathbf{x}}}{e^{-i2\alpha}-e^{-i2\beta}}\right),\\ \mathbf{x}_{\tilde{b}}&=-\frac{1}{A_{b}}\left(\frac{e^{i\beta}\mathbf{x}}{e^{i2\alpha}-e^{i2\beta}}+\frac{e^{-i\beta}\bar{\mathbf{x}}}{e^{-i2\alpha}-e^{-i2\beta}}\right)\end{split} (26)

are the new spatial bases (eigenvectors) associated with a~\widetilde{a} and b~\widetilde{b}, which are real-valued.

Proof.

Multiplying (24a) by AbeiαA_{b}e^{i\alpha} and multiplying (24b) by AaeiβA_{a}e^{i\beta} lead to

Abeiαa~=AaAbei2αv+AaAbv¯,Aaeiβb~=AaAbei2βv+AaAbv¯.\begin{split}A_{b}e^{i\alpha}\widetilde{a}&=A_{a}A_{b}e^{i2\alpha}v+A_{a}A_{b}\bar{v},\\ A_{a}e^{i\beta}\widetilde{b}&=A_{a}A_{b}e^{i2\beta}v+A_{a}A_{b}\bar{v}.\end{split}

Taking the difference between the above two equations yields

v=1AaAbAbeiαa~Aaeiβb~ei2αei2β,andv¯=1AaAbAbeiαa~Aaeiβb~ei2αei2β.v=\frac{1}{A_{a}A_{b}}\frac{A_{b}e^{i\alpha}\widetilde{a}-A_{a}e^{i\beta}\widetilde{b}}{e^{i2\alpha}-e^{i2\beta}},\qquad\mbox{and}\qquad\bar{v}=\frac{1}{A_{a}A_{b}}\frac{A_{b}e^{-i\alpha}\widetilde{a}-A_{a}e^{-i\beta}\widetilde{b}}{e^{-i2\alpha}-e^{-i2\beta}}. (27)

Recall in (23) that the full spatiotemporal evolution of the solution is given by 𝐮~=v𝐱+v¯𝐱¯\widetilde{\mathbf{u}}=v\mathbf{x}+\bar{v}\bar{\mathbf{x}}. Plugging (27) into (23) gives

𝐮~=v𝐱+v¯𝐱¯=1AaAbAbeiαa~Aaeiβb~ei2αei2β𝐱+1AaAbAbeiαa~Aaeiβb~ei2αei2β𝐱¯=1Aa(eiα𝐱ei2αei2β+eiα𝐱¯ei2αei2β)a~1Ab(eiβ𝐱ei2αei2β+eiβ𝐱¯ei2αei2β)b~:=𝐱a~a~+𝐱b~b~.\begin{split}{\widetilde{\mathbf{u}}}&={v\mathbf{x}+\bar{v}\bar{\mathbf{x}}}\\ &=\frac{1}{A_{a}A_{b}}\frac{A_{b}e^{i\alpha}\widetilde{a}-A_{a}e^{i\beta}\widetilde{b}}{e^{i2\alpha}-e^{i2\beta}}\mathbf{x}+\frac{1}{A_{a}A_{b}}\frac{A_{b}e^{-i\alpha}\widetilde{a}-A_{a}e^{-i\beta}\widetilde{b}}{e^{-i2\alpha}-e^{-i2\beta}}\bar{\mathbf{x}}\\ &=\frac{1}{A_{a}}\left(\frac{e^{i\alpha}\mathbf{x}}{e^{i2\alpha}-e^{i2\beta}}+\frac{e^{-i\alpha}\bar{\mathbf{x}}}{e^{-i2\alpha}-e^{-i2\beta}}\right)\widetilde{a}\\ &\quad\ -\frac{1}{A_{b}}\left(\frac{e^{i\beta}\mathbf{x}}{e^{i2\alpha}-e^{i2\beta}}+\frac{e^{-i\beta}\bar{\mathbf{x}}}{e^{-i2\alpha}-e^{-i2\beta}}\right)\widetilde{b}\\ {:}&{=\mathbf{x}_{\tilde{a}}\widetilde{a}+\mathbf{x}_{\tilde{b}}\widetilde{b}}.\end{split} (28)

The above framework is adapted to the cases where the main interest lies in certain averaged quantities, as in many of the classical conceptual models for the ENSO. For example, a~\widetilde{a} can be the SST in the eastern Pacific and b~\widetilde{b} the thermocline in the western Pacific. They can also be the atmosphere wind and ocean current averaged over the entire Pacific or a specific region. In such situations, the amplitudes AaA_{a} and AbA_{b} can be determined by two scalars xa~x_{\tilde{a}} and xb~x_{\tilde{b}}, which are the averaged values of 𝐱a~\mathbf{x}_{\tilde{a}} and 𝐱b~\mathbf{x}_{\tilde{b}} over a certain domain. This allows the development of a two-dimensional system for a~\widetilde{a} and b~\widetilde{b} that is linked with the traditional conceptual models.

Proposition 5.3.

Let xa:=Aaeiαx_{a}:=A_{a}e^{i\alpha} and xb:=Abeiβx_{b}:=A_{b}e^{i\beta} such that in (24) a~=xav+x¯av¯\widetilde{a}=x_{a}v+\bar{x}_{a}\bar{v} and b~=xbv+x¯bv¯\widetilde{b}=x_{b}v+\bar{x}_{b}\bar{v}. Then the two-dimensional conceptual model for a~\widetilde{a} and b~\widetilde{b} yields

da~dt=doa~+c~11a~+c~12b~+2ga,Re,db~dt=dob~+c~21a~+c~22b~+2gb,Re,\begin{split}\frac{{\,\rm d}\widetilde{a}}{{\,\rm d}t}&=-d_{o}\widetilde{a}+\widetilde{c}_{11}\widetilde{a}+\widetilde{c}_{12}\widetilde{b}+2g_{a,Re},\\ \frac{{\,\rm d}\widetilde{b}}{{\,\rm d}t}&=-d_{o}\widetilde{b}+\widetilde{c}_{21}\widetilde{a}+\widetilde{c}_{22}\widetilde{b}+2g_{b,Re},\end{split} (29)

where all the coefficients are real. In (29),

C~=(c~11c~12c~21c~22)=ωo1cRedRe(cRedImcImdImdRecIm).\widetilde{C}=\left(\begin{array}[]{cc}\widetilde{c}_{11}&\widetilde{c}_{12}\\ \widetilde{c}_{21}&\widetilde{c}_{22}\\ \end{array}\right)=-\frac{\omega_{o}}{1-c_{Re}d_{Re}}\left(\begin{array}[]{cc}c_{Re}d_{Im}&c_{Im}\\ d_{Im}&d_{Re}c_{Im}\\ \end{array}\right). (30)

with c=cRe+cIm:=xaxbc=c_{Re}+c_{Im}:=\frac{x_{a}}{x_{b}} and d=dRe+dIm:=xbxad=d_{Re}+d_{Im}:=\frac{x_{b}}{x_{a}}. The two forcing terms are given by ga,Re=Re(gxa)=Re(𝐳T𝐬𝐮xa)apg_{a,Re}=Re(gx_{a})=Re(\mathbf{z}^{T}\mathbf{s}_{\mathbf{u}}x_{a})a_{p} and gb,Re=Re(gxb)=Re(𝐳T𝐬𝐮xb)apg_{b,Re}=Re(gx_{b})=Re(\mathbf{z}^{T}\mathbf{s}_{\mathbf{u}}x_{b})a_{p}.

See the Appendix for the proofs.

5.3 Special case: stochastic discharge-recharge oscillator

The discharge-recharge oscillator argues that discharge and recharge of equatorial heat content cause the coupled system to oscillate [7]. Define a~:=TE\widetilde{a}:=T_{E} and b~:=HW\tilde{b}:=H_{W} in (29), representing the averaged SST in the eastern Pacific and the averaged thermocline depth in the western Pacific. The corresponding scalars xTEx_{T_{E}} and xHWx_{H_{W}} are given by

xTE=0.1302+0.0106iandxHW=0.0279+0.0375i,x_{T_{E}}=0.1302+0.0106i\qquad\mbox{and}\qquad x_{H_{W}}=-0.0279+0.0375i, (31)

which are computed by averaging the eigenvectors associated with TT and HH over the eastern Pacific and western Pacific, respectively. See Figure 8.

Refer to caption
Figure 8: The eigenvectors of vv associated with TT and HH (top) and their averaged values over the eastern and western Pacific (bottom).

The other coefficients in (29) can be computed using the eigenvalues of the leading two modes λo\lambda_{o} as well as xTEx_{T_{E}} and xHWx_{H_{W}} in (31). They are:

C~=(c~11c~12c~21c~22)=(0.07520.39650.05090.0752)\widetilde{C}=\left(\begin{array}[]{cc}\widetilde{c}_{11}&\widetilde{c}_{12}\\ \widetilde{c}_{21}&\widetilde{c}_{22}\end{array}\right)=\left(\begin{array}[]{cc}0.0752&0.3965\\ -0.0509&-0.0752\end{array}\right) (32)

and

αTE=2Re(𝐳T𝐬𝐮xTE)=1.0094,αHW=2Re(𝐳T𝐬𝐮xHW)=0.4217.\alpha_{T_{E}}=2\operatorname{Re}\left(\mathbf{z}^{T}\mathbf{s}_{\mathbf{u}}x_{T_{E}}\right)=1.0094,\quad\alpha_{H_{W}}=2\operatorname{Re}\left(\mathbf{z}^{T}\mathbf{s}_{\mathbf{u}}x_{H_{W}}\right)=-0.4217. (33)

The remaining thing is to determine the multiplicative noise coefficient σp(TW)\sigma_{p}(T_{W}) of the stochastic amplitude apa_{p} in (9), which is a function of the averaged SST in the western Pacific TWT_{W}. Yet, TWT_{W} is not the state variable in the two-dimensional discharge-recharge oscillator model. Nevertheless, TWT_{W} can be written as a linear function of TET_{E} and HWH_{W}. Recall that

TE=xTEv+x¯TEv¯andHW=xHWv+x¯HWv¯,T_{E}=x_{T_{E}}v+\bar{x}_{T_{E}}\bar{v}\qquad\mbox{and}\qquad H_{W}=x_{H_{W}}v+\bar{x}_{H_{W}}\bar{v}, (34)

which can be used to solve Re(v)\mbox{Re}(v) and Im(v)\mbox{Im}(v),

Re(v)=TEIm(xHW)HWIm(xTE)2(Re(xTE)Im(xHW)Im(xTE)Re(xHW)),Im(v)=TERe(xHW)HWRe(xTE)2(Re(xTE)Im(xHW)Im(xTE)Re(xHW)).\begin{split}\mbox{Re}(v)&=\frac{T_{E}\mbox{Im}(x_{H_{W}})-H_{W}\mbox{Im}(x_{T_{E}})}{2(\mbox{Re}(x_{T_{E}})\mbox{Im}(x_{H_{W}})-\mbox{Im}(x_{T_{E}})\mbox{Re}(x_{H_{W}}))},\\ \mbox{Im}(v)&=\frac{T_{E}\mbox{Re}(x_{H_{W}})-H_{W}\mbox{Re}(x_{T_{E}})}{2(\mbox{Re}(x_{T_{E}})\mbox{Im}(x_{H_{W}})-\mbox{Im}(x_{T_{E}})\mbox{Re}(x_{H_{W}}))}.\\ \end{split} (35)

Once the basis vv is determined, the value of TWT_{W} can be computed,

TW=xTWv+x¯TWv¯T_{W}=x_{T_{W}}v+\bar{x}_{T_{W}}\bar{v} (36)

which are functions of HWH_{W} and TET_{E}, and xTWx_{T_{W}} is given by

xTW=0.0272+0.0484i.x_{T_{W}}=-0.0272+0.0484i. (37)

The direct representation of the TWT_{W} by TET_{E} also justifies and unifies the use of SST in different regions (averaged values over the eastern Pacific, over the western Pacific or basin averaged value) as multiplicative noise coefficients in various conceptual models [58, 59, 60, 61].

After obtaining the value of TWT_{W}, the multiplicative noise coefficient σp(TW)\sigma_{p}(T_{W}) is determined. Note that the σp(TW)\sigma_{p}(T_{W}) is slightly modified to σp0=0.2\sigma_{p0}=0.2 and σp1=1.8\sigma_{p1}=1.8 to compensate for the approximation error from the low-order representation.

According to (32), the coefficient c~11\widetilde{c}_{11} and c~12\widetilde{c}_{12} associated with TET_{E} equations are both positive. The coefficient c~11>0\widetilde{c}_{11}>0 represents the Bjerknes positive feedback while c~12>0\widetilde{c}_{12}>0 is the recharge mechanism. On the other hand, a negative coefficient c~21\widetilde{c}_{21} appears in the equation of HWH_{W}, which indicates the discharge mechanism. Another finding is that, although the wind bursts directly impact the SST in heuristic thinking, the stochastic forcing appears in both the equations of TET_{E} and HWH_{W} for the completeness of the stochastic discharge-recharge diagram. The opposite signs of αTE\alpha_{T_{E}} and αHW\alpha_{H_{W}} imply that the increase of SST in the eastern Pacific by the WWBs will simultaneously cause a decrease of the thermocline in the western Pacific.

Figure 9 shows the time series and phase plot of both the deterministic and stochastic versions of the discharge-recharge models developed above. The deterministic model (by setting the damping to be zero and ignoring the stochastic terms) leads to a regular oscillator while the time series of HWH_{W} and TET_{E} generated from the stochastic model mimic those in observations. It is worth highlighting that the phase difference between HWH_{W} and TET_{E} is 122o122^{o} (not π/2\pi/2 or 90o90^{o}), which is also confirmed by the lag between the two time series. Such a phase difference also resembles that from the observations. In addition, the observational time series and the stochastic model simulation of TET_{E} are compared. Due to the random noise, the trajectories from the stochastic model and the observations do not expect to have a one-to-one point-wise match between each other. Nevertheless, the stochastic model simulation succeeds in capturing the EP events. Finally, the PDF and the autocorrelation function (ACF) of TET_{E} from the stochastic model have very similar behavior to the observations. The PDF is skewed with a one-sided fat tail, which indicates that the model can reproduce a sufficient number of the super El Niño events. Similarly, matching the ACF with observations implies that the time series of TET_{E} from the model has, on average, the same temporal correlation (namely the time scale of the memory effect) as the observations. These findings suggest that the low-order model can capture the observed dynamical and statistical features of the time series TET_{E}, which is an important indicator for the eastern Pacific El Niño.

Refer to caption
Figure 9: Comparison of the time series and the phase plot between the deterministic and the stochastic versions of the discharge-recharge model; the time series, PDFs, and ACFs of TET_{E} between the stochastic version and the observations. Panel(a): The time series of TET_{E} and HWH_{W} generated from two versions of the discharge-recharge model. Panel (b): The phase plot of TET_{E} and HWH_{W} from two versions of the model, and observations. Here, the HWH_{W} in two version models are multiplied by a constant to keep the same standard deviation of the observations. Panel (c): The time series of TET_{E} from the stochastic version of the model and observations. Panel (d): The PDFs of TET_{E} for the solution to the stochastic version of the model and observations. Panel (e): The ACFs of TET_{E} from the stochastic version of the model and observations. In Panel (d)-(e), red and blue curves are for the observation and stochastic model, respectively. For the stochastic model, the total 20000-year-long simulation is divided into 500 non-overlapping segments, each of which has a 40-year period as the observation. Then the average (blue line) and its one standard deviation intervals (shading) are illustrated.

5.4 Special case: stochastic delayed oscillator

The two-dimensional stochastic oscillator (29) can be further reduced to a one-dimensional oscillator with a delayed effect, which leads to the delayed oscillator of the ENSO.

Proposition 5.4.

Defining a~:=TE\widetilde{a}:=T_{E}, the averaged SST over the eastern Pacific. The delayed oscillator is given by

dTE(t)dt=doTE(t)+c~11TE(t)+c~12cyedoφ/ωoTE(tφ/ωo)+αTEap(t)+2cyc~12tφ/ωotedo(ts)[y1,Re(cos(ω)αTE,1sin(ω)αTE,2)y1,Im(cos(ω)αTE,2+sin(ω)αTE,1)]ap(s)ds\begin{split}\frac{{\,\rm d}T_{E}(t)}{{\,\rm d}t}&=-d_{o}T_{E}(t)+\widetilde{c}_{11}T_{E}(t)+\widetilde{c}_{12}c_{y}e^{-d_{o}\varphi/\omega_{o}}T_{E}(t-\varphi/\omega_{o})+\alpha_{T_{E}}a_{p}(t)\\ &\qquad+2c_{y}\widetilde{c}_{12}\int_{t-\varphi/\omega_{o}}^{t}e^{-d_{o}(t-s)}\bigg{[}y_{1,Re}\Big{(}\cos(\omega)\alpha_{T_{E},1}-\sin(\omega)\alpha_{T_{E},2}\Big{)}\\ &\qquad\qquad\qquad-y_{1,Im}\Big{(}\cos(\omega)\alpha_{T_{E},2}+\sin(\omega)\alpha_{T_{E},1}\Big{)}\bigg{]}a_{p}(s){\,\rm d}s\end{split} (38)

where φ:=αβ\varphi:=\alpha-\beta and ω=ωo(ts)φ\omega=\omega_{o}(t-s)-\varphi. The coefficients c~11\widetilde{c}_{11}, c~12\widetilde{c}_{12} and αTE\alpha_{T_{E}} are the same as those in (32) and (33). The other two noise coefficients αTE,1\alpha_{T_{E},1} and αTE,2\alpha_{T_{E},2} are given by

αTE,1=Re(r12Re(𝐳T𝐬𝐮xTE)+r22Re(𝐳T𝐬𝐮xHW)),αTE,2=Im(r12Re(𝐳T𝐬𝐮xTE)+r22Re(𝐳T𝐬𝐮xHW)),\begin{split}\alpha_{T_{E},1}&=Re\Big{(}r_{1}2Re(\mathbf{z}^{T}\mathbf{s}_{\mathbf{u}}x_{T_{E}})+r_{2}2Re(\mathbf{z}^{T}\mathbf{s}_{\mathbf{u}}x_{H_{W}})\Big{)},\\ \alpha_{T_{E},2}&=Im\Big{(}r_{1}2Re(\mathbf{z}^{T}\mathbf{s}_{\mathbf{u}}x_{T_{E}})+r_{2}2Re(\mathbf{z}^{T}\mathbf{s}_{\mathbf{u}}x_{H_{W}})\Big{)},\end{split}

Here, r1r_{1} and r2r_{2} are solved as follows. As the matrix C~\widetilde{C} in (30) has two distinct eigenvalues, which are also complex conjugates, it can be diagonalized C~=𝐘Λ𝐘1\widetilde{C}=\mathbf{Y}\Lambda\mathbf{Y}^{-1}. Denote by

𝐘=(y1y¯1y2y¯2)and𝐘1=(r1r2r¯1r¯2)\mathbf{Y}=\left(\begin{array}[]{cc}y_{1}&\bar{y}_{1}\\ y_{2}&\bar{y}_{2}\\ \end{array}\right)\qquad\mbox{and}\qquad\mathbf{Y}^{-1}=\left(\begin{array}[]{cc}r_{1}&r_{2}\\ \bar{r}_{1}&\bar{r}_{2}\\ \end{array}\right)

where the fact that the two eigenvectors are complex conjugates has been utilized. In addition, the angle difference φ\varphi and the amplitude difference cyc_{y} are given by

My2=cyMy1Φ,M_{y_{2}}=c_{y}M_{y_{1}}\varPhi, (39)

where

My1=(y1,Rey1,Imy1,Imy1,Re)My2=(y2,Rey2,Imy2,Imy2,Re),Φ=(cosφsinφsinφcosφ),cy=y2y1,\begin{split}M_{y_{1}}=\left(\begin{array}[]{cc}y_{1,Re}&-y_{1,Im}\\ -y_{1,Im}&-y_{1,Re}\\ \end{array}\right)\qquad&M_{y_{2}}=\left(\begin{array}[]{cc}y_{2,Re}&-y_{2,Im}\\ -y_{2,Im}&-y_{2,Re}\\ \end{array}\right),\\ \varPhi=\left(\begin{array}[]{cc}\cos\varphi&\sin\varphi\\ -\sin\varphi&\cos\varphi\\ \end{array}\right),\qquad&c_{y}=\frac{\|y_{2}\|}{\|y_{1}\|},\end{split}

with y1=y1,Re+iy1,Imy_{1}=y_{1,Re}+iy_{1,Im} and y2=y2,Re+iy2,Imy_{2}=y_{2,Re}+iy_{2,Im}.

See the Appendix for the proofs. Numerically, these constants and coefficients are given by

y1=0.9414,y2=0.1785i0.2861,r1=0.5311+i0.3314,r2=i1.7476,αTE=1.0094,αTE,1=0.5361,αTE,2=0.4024,φ=2.1287,cy=0.3582.\begin{gathered}y_{1}=0.9414,\ y_{2}=-0.1785-i0.2861,\ r_{1}=0.5311+i0.3314,\ r_{2}=i1.7476,\\ \alpha_{T_{E}}=1.0094,\ \alpha_{T_{E},1}=0.5361,\ \alpha_{T_{E},2}=0.4024,\ \varphi=2.1287,\ c_{y}=0.3582.\end{gathered}

The detailed expression of the stochastic forcing with the memory effect derived here is an essential supplement to the original delayed oscillator model [29, 16], which was deterministic. Several conclusions can be drawn from this delayed oscillator. First, In front of the delayed term TE(tφ/ωo)T_{E}\left(t-\varphi/\omega_{o}\right), there is a factor with an exponential decay edoφ/ωoe^{-d_{o}\varphi/\omega_{o}}. This is a crucial feature, which indicates that the contribution of the delayed signal will be affected by the damping effect and becomes weaker as time goes on. Second, there are two ways that stochastic noise contributes to the system. The term αTEap(t)\alpha_{T_{E}}a_{p}(t) provides an instantaneous response to the system, while the stochastic forcing inside the integral represents the contribution from the noise accumulation. The latter also has a delayed contribution. Notably, although the damping time of the noise dpd_{p} is intraseasonal, the noise interacting with the coupled system can affect the interannual time scale. Third, similar to the discharge-recharge oscillator, the delay here is φ=2.1287\varphi=2.1287 (122o122^{o}) instead of π/2\pi/2 (90o90^{o}). Finally, the sign of the delayed term depends on the sign of c~12\widetilde{c}_{12}, which can be either positive or negative, depending on the state variables used for the model. The parameter c~12\widetilde{c}_{12} is positive in the above model since φ=2.1287>π/2\varphi=2.1287>\pi/2. Yet, when the angle φ\varphi is less than π/2\pi/2, then c~12<0\widetilde{c}_{12}<0 representing a negative delay, which is consistent with the discussions in [16]. In particular, if the average of the thermocline over the whole Pacific is adopted as was in [16], then a negative delay (about π/6\pi/6) appears.

6 Conclusion and Discussions

In this paper, a framework of low-order stochastic conceptual models for the ENSO is systematically derived from a spatially-extended stochastic dynamical system with full mathematical rigor. It provides outputs for all the atmosphere, ocean, and SST components with detailed spatiotemporal patterns, which differs from many existing conceptual models focusing only on a small set of specified state variables. The framework also allows rigorously deriving the stochastic versions of the discharge-recharge oscillator and delayed oscillator models. Observational data is utilized to validate the skill of the low-order conceptual model in capturing the large-scale features of the ENSO in the eastern Pacific. The rigorous derivation of these low-order models provides a unique connection between models with different complexities. The framework developed here also facilitates understanding the instantaneous and memory effects of stochastic noise in contributing to the large-scale dynamics of ENSO.

The current modeling framework focuses on the eastern Pacific ENSO events. To characterize the ENSO diversity [55] that also includes the central Pacific El Niño, the ocean zonal advection mechanism should be incorporated in the starting spatially-extended model. It is worth noting that the dynamics of the central Pacific El Niño often contain certain levels of nonlinearity [62, 9]. Therefore, suitable nonlinearity needs to be added to the linear low-order conceptual models to better describe the central Pacific El Niño, which is remained as a future work.

Despite the simple forms of conceptual models, they can be utilized as testbeds for guiding more sophisticated models. In particular, including stochastic noise or stochastic parameterizations in ICMs or GCMs is always a challenging topic. Therefore, testing various strategies based on conceptual models is practically useful. These simple conceptual models can also be utilized to build a hierarchy of models with different complexity for understanding the ENSO dynamics with the help of data-driven techniques.

Acknowledgements

The research of N.C. is partially funded by ONR N00014-21-1-2904. Y.Z. is supported as a PhD research assistant under this grant. N.C. wish to thank his late advisor Andrew J. Majda for valuable discussions on some initial ideas of this work.

7 Appendix

7.1 A summary of the parameters in the spatially-extended stochastic system

This section will provide details for parameter values and dimensional values.

Variable Explanation Value
HH thermocline depth anomalies 20.8m20.8\mathrm{~{}m}
TT sea surface temperature anomalies 1.5K1.5\mathrm{~{}K}
UU zonal current speed anomalies 0.25m/s0.25\mathrm{~{}m}/\mathrm{s}
tt time axis intraseasonal 0.48 days
uu zonal wind speed anomalies 5m/s5\mathrm{~{}m}/\mathrm{s}
θ\theta potential temperature anomalies 1.5K1.5\mathrm{~{}K}
Table 1: Physical dimensions of model parameters.
Parameter Explanation Value
χA\chi_{A} atmospheric meridional projection coefficient 0.325
χO\chi_{O} oceanic meridional projection coefficient 1.3
LAL_{A} equatorial belt length 8/3
LOL_{O} equatorial Pacific length 7/6
Q¯\bar{Q} mean vertical moisture gradient 0.9
T¯\bar{T} mean SST 16.67 (which is 25oC)
γ\gamma wind stress coefficient 6.53
ζ\zeta latent heating exchange coefficient 8.7
αq\alpha_{q} latent heating factor αq=qcqeexp(qeT¯)/τq\alpha_{q}=q_{c}q_{e}\exp(q_{e}\bar{T})/\tau_{q}
qcq_{c} latent heating multiplier coefficient 7
qeq_{e} latent heating exponential coefficient 0.093
τq\tau_{q} latent heating adjustment rate 15
rWr_{W} western boundary reflection coefficient 0.5
rEr_{E} eastern boundary reflection coefficient 1
c1c_{1} modified ratio of phase speed 0.5
η\eta profile of thermocline feedback η(x)=1.5+tanh(7.5(xLO/2))/2\eta(x)=1.5+\tanh(7.5(x-L_{O}/2))/2
dAd_{A} additional damping 10810^{-8}
dpd_{p} wind burst damping 3.4 (which is 3mon-1)
sps_{p} wind burst zonal structure sp(x)=exp(45(xLO/4)2)s_{p}(x)=\exp(-45(x-L_{O}/4)^{2})
σp0\sigma_{p0} wind burst source of quiescent state 0.2
σp1\sigma_{p1} wind burst source of active state 2.6
μ10\mu_{10} transition rate from the active to quiescent state μ10=(1tanh(2TW))/4\mu_{10}=\left(1-\tanh\left(2T_{W}\right)\right)/4
μ01\mu_{01} transition rate from the quiescent to active state μ01=(tanh(2TW)+1)/8\mu_{01}=\left(\tanh\left(2T_{W}\right)+1\right)/8
Table 2: Model parameter values.

7.2 Determining the matrix 𝐌\mathbf{M} in (14)

Recall the system in (14). Rewrite it in terms of the components with respect to the three set of the variables 𝐊𝐎\mathbf{K_{O}}, 𝐑𝐎\mathbf{R_{O}} and 𝐓\mathbf{T}:

ddt(𝐊𝐎𝐑𝐎𝐓)=(𝐌11𝐌12𝐌13𝐌21𝐌22𝐌23𝐌31𝐌32𝐌33)(𝐊𝐎𝐑𝐎𝐓),\frac{{\,\rm d}}{{\,\rm d}t}\left(\begin{array}[]{c}\mathbf{K_{O}}\\ \mathbf{R_{O}}\\ \mathbf{T}\\ \end{array}\right)=\left(\begin{array}[]{ccc}\mathbf{M}_{11}&\mathbf{M}_{12}&\mathbf{M}_{13}\\ \mathbf{M}_{21}&\mathbf{M}_{22}&\mathbf{M}_{23}\\ \mathbf{M}_{31}&\mathbf{M}_{32}&\mathbf{M}_{33}\\ \end{array}\right)\left(\begin{array}[]{c}\mathbf{K_{O}}\\ \mathbf{R_{O}}\\ \mathbf{T}\\ \end{array}\right), (40)

where the matrix 𝐌\mathbf{M} in (14) is expanded into a 3×33\times 3 matrix. In the following, the goal is to determine the matrix 𝐌\mathbf{M}, namely the 𝐌ij\mathbf{M}_{ij} with i,j=1,2,3i,j=1,2,3, in (40).

Let us start with the ocean model

tKO+c1xKO\displaystyle\partial_{t}K_{O}+c_{1}\partial_{x}K_{O} =a2(KARA),\displaystyle=\frac{a}{2}\ (K_{A}-R_{A}),\qquad\qquad KO(0,t)=rWRO(0,t),\displaystyle K_{O}(0,t)=r_{W}R_{O}(0,t), (41)
tROc13xRO\displaystyle\partial_{t}R_{O}-\frac{c_{1}}{3}\partial_{x}R_{O} =a3(KARA),\displaystyle=-\frac{a}{3}\ (K_{A}-R_{A}),\qquad\qquad RO(LO,t)=rEKO(LO,t),\displaystyle R_{O}(L_{O},t)=r_{E}K_{O}(L_{O},t),

An upwind scheme is utilized for the spatial discretization,

tKO,i=c1Δx(KO,iKO,i1)+a2(KA,iRA,i),i=1,,NO,tRO,i=c13Δx(RO,i+1RO,i)a3(KA,iRA,i),i=1,,NO,\begin{split}\partial_{t}K_{O,i}&=-\frac{c_{1}}{\Delta x}\left(K_{O,i}-K_{O,{i-1}}\right)+\frac{a}{2}\left(K_{A,i}-R_{A,i}\right),\qquad i=1,\ldots,N_{O},\\ \partial_{t}R_{O,i}&=\frac{c_{1}}{3\Delta x}\left(R_{O,{i+1}}-R_{O,i}\right)-\frac{a}{3}\left(K_{A,i}-R_{A,i}\right),\qquad i=1,\ldots,N_{O},\end{split} (42)

where the boundary conditions are given by KO,0=rWRO,1K_{O,0}=r_{W}R_{O,1} and RO,NO+1=rEKO,NOR_{O,{N_{O}+1}}=r_{E}K_{O,{N_{O}}}. Therefore,

𝐌11=c1Δx(111111111)NO×NO,𝐌12=c1Δx(rW)NO×NO𝐌22=c13Δx(111111111)NO×NO,𝐌21=c13Δx(rE)NO×NO{\begin{gathered}\mathbf{M}_{11}=-\frac{c_{1}}{\Delta x}\left(\begin{array}[]{cccccc}1&&&&&\\ -1&1&&&&\\ &-1&1&&&\\ &&\ddots&\ddots&&\\ &&&-1&1&\\ &&&&-1&1\\ \end{array}\right)_{N_{O}\times N_{O}},\ \mathbf{M}_{12}=-\frac{c_{1}}{\Delta x}\left(\begin{array}[]{cccccc}-r_{W}&&&&&\\ &&&&&\\ &&&&&\\ &&&&&\\ &&&&&\\ &&&&&\\ \end{array}\right)_{N_{O}\times N_{O}}\\ \mathbf{M}_{22}=\frac{c_{1}}{3\Delta x}\left(\begin{array}[]{cccccc}-1&1&&&&\\ &-1&1&&&\\ &&\ddots&\ddots&&\\ &&&-1&1&\\ &&&&-1&1\\ &&&&&-1\\ \end{array}\right)_{N_{O}\times N_{O}},\quad\mathbf{M}_{21}=\frac{c_{1}}{3\Delta x}\left(\begin{array}[]{cccccc}&&&&&\\ &&&&&\\ &&&&&\\ &&&&&\\ &&&&&\\ &&&&&r_{E}\\ \end{array}\right)_{N_{O}\times N_{O}}\end{gathered}} (43)

Next, for the atmosphere model

dAKA+xKA\displaystyle d_{A}K_{A}+\partial_{x}K_{A} =m1(𝟏[0,LO]αqTEq)/αq,\displaystyle=m_{1}(\mathbf{1}_{[0,L_{O}]}\alpha_{q}T-\langle E_{q}\rangle)/\alpha_{q},\qquad\qquad KA(0,t)=KA(LA,t),\displaystyle K_{A}(0,t)=K_{A}(L_{A},t), (44)
dARAxRA/3\displaystyle d_{A}R_{A}-\partial_{x}R_{A}/3 =m2(𝟏[0,LO]αqTEq)/αq,\displaystyle=m_{2}(\mathbf{1}_{[0,L_{O}]}\alpha_{q}T-\langle E_{q}\rangle)/\alpha_{q},\qquad\qquad RA(0,t)=RA(LA,t),\displaystyle R_{A}(0,t)=R_{A}(L_{A},t),

where 𝟏[0,LO]\mathbf{1}_{[0,L_{O}]} is an indicator function with value being 11 when variables are located within the interval [0,LO][0,L_{O}] and 0 otherwise. The discretization of (44) results in

dAKA,i+1Δx(KA,i+1KA,i)=m1(Ti1NAj=1NOTj),dARA,i13Δx(RA,i+1RA,i)=m2(Ti1NAj=1NOTj),i=1,,NO,{\begin{split}d_{A}K_{A,i}+\frac{1}{\Delta x}(K_{A,{i+1}}-K_{A,{i}})&=m_{1}\left(T_{i}-\frac{1}{N_{A}}\sum_{j=1}^{N_{O}}T_{j}\right),\\ d_{A}R_{A,i}-\frac{1}{3\Delta x}(R_{A,{i+1}}-R_{A,{i}})&=m_{2}\left(T_{i}-\frac{1}{N_{A}}\sum_{j=1}^{N_{O}}T_{j}\right),\qquad i=1,\ldots,N_{O},\end{split}}

and

dAKA,i+1Δx(KA,i+1KA,i)=m1(1NAj=1NOTj),dARA,i13Δx(RA,i+1RA,i)=m2(1NAj=1NOTj),i=NO+1,,NA,{\begin{split}d_{A}K_{A,i}+\frac{1}{\Delta x}(K_{A,{i+1}}-K_{A,{i}})&=m_{1}\left(-\frac{1}{N_{A}}\sum_{j=1}^{N_{O}}T_{j}\right),\\ d_{A}R_{A,i}-\frac{1}{3\Delta x}(R_{A,{i+1}}-R_{A,{i}})&=m_{2}\left(-\frac{1}{N_{A}}\sum_{j=1}^{N_{O}}T_{j}\right),\qquad i=N_{O}+1,\ldots,N_{A},\end{split}} (45)

where the relationship Eq=αqTE_{q}=\alpha_{q}T has been used. Note that the averaging is taken over the entire equator band [0,NA][0,N_{A}] and therefore the factor 1/NA1/N_{A} in front of the summation appears. Rearranging the terms in (45) gives

KA,i+1+(dAΔx1)KA,i=NA1NAΔxm1(Ti1NA11jNOjiTj),RA,i+1+(3dAΔx+1)RA,i=3NA1NAΔxm2(Ti1NA11jNOjiTj),i=1,NO{\begin{split}K_{A,{i+1}}+(d_{A}\Delta x-1)K_{A,{i}}&=\frac{N_{A}-1}{N_{A}}\Delta xm_{1}\left(T_{i}-\frac{1}{N_{A}-1}\sum_{1\leq j\leq N_{O}}^{j\neq i}T_{j}\right),\\ -R_{A,{i+1}}+(3d_{A}\Delta x+1)R_{A,i}&=3\frac{N_{A}-1}{N_{A}}\Delta xm_{2}\left(T_{i}-\frac{1}{N_{A}-1}\sum_{1\leq j\leq N_{O}}^{j\neq i}T_{j}\right),\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad i=1,\ldots N_{O}\end{split}}

and

KA,i+1+(dAΔx1)KA,i=NA1NAΔxm1(1NA11jNOTj),RA,i+1+(3dAΔx+1)RA,i=3NA1NAΔxm2(1NA11jNOTj),i=NO+1,,NA{\begin{split}K_{A,{i+1}}+(d_{A}\Delta x-1)K_{A,{i}}&=\frac{N_{A}-1}{N_{A}}\Delta xm_{1}\left(-\frac{1}{N_{A}-1}\sum_{1\leq j\leq N_{O}}T_{j}\right),\\ -R_{A,{i+1}}+(3d_{A}\Delta x+1)R_{A,i}&=3\frac{N_{A}-1}{N_{A}}\Delta xm_{2}\left(-\frac{1}{N_{A}-1}\sum_{1\leq j\leq N_{O}}T_{j}\right),\\ &\qquad\qquad\qquad\qquad\qquad\ i=N_{O}+1,\ldots,N_{A}\end{split}}

Thus, KA,iK_{A,i} and RA,iR_{A,i} for i=1,NOi=1,\ldots N_{O} can be solved via the following linear systems,

𝐌𝐊𝐊𝐀=C~1𝐁𝐊𝐓,𝐌𝐑𝐑𝐀=C~2𝐁𝐑𝐓,\begin{split}\mathbf{M_{K}}\cdot\mathbf{K_{A}}&=\widetilde{C}_{1}\mathbf{B_{K}}\mathbf{T},\\ \mathbf{M_{R}}\cdot\mathbf{R_{A}}&=\widetilde{C}_{2}\mathbf{B_{R}}\mathbf{T},\end{split} (46)

which means 𝐊𝐀\mathbf{K_{A}} and 𝐑𝐀\mathbf{R_{A}} can be expressed by 𝐓\mathbf{T}. In (46)

C~1=NA1NAΔxm1andC~2=3NA1NAΔxm2\widetilde{C}_{1}=\frac{N_{A}-1}{N_{A}}\Delta xm_{1}\qquad\mbox{and}\qquad\widetilde{C}_{2}=3\frac{N_{A}-1}{N_{A}}\Delta xm_{2}

with Δx\Delta{x} being one unit of the spatial discretization. Further define κ=dAΔx\kappa=d_{A}\Delta x. Then the matrices in (46) are given by

𝐌𝐊=(1+κ11+κ11+κ11+κ111+κ)NA×NA,𝐊𝐀=(KA,1KA,2KA,NOKA,NA1KA,NA)NA×1,{\mathbf{M_{K}}=\left(\begin{array}[]{cccccccc}-1+\kappa&1&&&&&\\ &-1+\kappa&1&&&&\\ &&\ddots&\ddots&&&\\ &&&-1+\kappa&1&&\\ &&&&\ddots&\ddots&\\ &&&&&-1+\kappa&1\\ 1&&&&&&-1+\kappa\\ \end{array}\right)_{N_{A}\times N_{A}},\ \mathbf{K_{A}}=\left(\begin{array}[]{c}K_{A,1}\\ K_{A,2}\\ \vdots\\ K_{A,{N_{O}}}\\ \vdots\\ K_{A,{N_{A}-1}}\\ K_{A,{N_{A}}}\\ \end{array}\right)_{N_{A}\times 1}},
𝐌𝐑=(1+3κ11+3κ11+3κ11+3κ111+3κ)NA×NA,𝐑𝐀=(RA,1RA,2RA,NORA,NA1RA,NA)NA×1.{\mathbf{M_{R}}=\left(\begin{array}[]{cccccccc}1+3\kappa&-1&&&&&\\ &1+3\kappa&-1&&&&\\ &&\ddots&\ddots&&&\\ &&&1+3\kappa&-1&&\\ &&&&\ddots&\ddots&\\ &&&&&1+3\kappa&-1\\ -1&&&&&&1+3\kappa\\ \end{array}\right)_{N_{A}\times N_{A}},\ \mathbf{R_{A}}=\left(\begin{array}[]{c}R_{A,1}\\ R_{A,2}\\ \vdots\\ R_{A,{N_{O}}}\\ \vdots\\ R_{A,{N_{A}-1}}\\ R_{A,{N_{A}}}\\ \end{array}\right)_{N_{A}\times 1}}.

and

𝐁𝐊=𝐁𝐑=(11NA11NA11NA111NA11NA11NA111NA11NA11NA11NA11NA11NA1)NA×NO,𝐓=(T1TNO)NO×1.{\mathbf{B}_{\mathbf{K}}=\mathbf{B_{R}}=\left(\begin{array}[]{ccccccc}1&-\frac{1}{N_{A}-1}&\ldots&-\frac{1}{N_{A}-1}\\ -\frac{1}{N_{A}-1}&1&\ldots&-\frac{1}{N_{A}-1}\\ \vdots&\vdots&\ddots&\vdots\\ -\frac{1}{N_{A}-1}&-\frac{1}{N_{A}-1}&\ldots&1\\ -\frac{1}{N_{A}-1}&-\frac{1}{N_{A}-1}&\ldots&-\frac{1}{N_{A}-1}\\ \vdots&\vdots&\ddots&\vdots\\ -\frac{1}{N_{A}-1}&-\frac{1}{N_{A}-1}&\ldots&-\frac{1}{N_{A}-1}\\ \end{array}\right)_{N_{A}\times N_{O}},\ \mathbf{T}=\left(\begin{array}[]{c}T_{1}\\ \vdots\\ T_{N_{O}}\\ \end{array}\right)_{N_{O}\times 1}}.

Recall the ocean model,

tKO,i=c1Δx(KO,iKO,i1)+a2(KA,iRA,i),tRO,i=c13Δx(RO,i+1RO,i)a3(KA,iRA,i),\begin{split}\partial_{t}K_{O,i}&=-\frac{c_{1}}{\Delta x}\left(K_{O,i}-K_{O,{i-1}}\right){+\frac{a}{2}\left(K_{A,i}-R_{A,i}\right)},\\ \partial_{t}R_{O,i}&=\frac{c_{1}}{3\Delta x}\left(R_{O,{i+1}}-R_{O,i}\right){-\frac{a}{3}\left(K_{A,i}-R_{A,i}\right)},\end{split}

Since 𝐊𝐀\mathbf{K_{A}} and 𝐑𝐀\mathbf{R_{A}} can be expressed by 𝐓\mathbf{T}, we have

𝐌13=a2[C~1𝐌𝐊1𝐁𝐊C~2𝐌𝐑1𝐁𝐑]|Row {1:NO},𝐌23=a3[C~1𝐌𝐊1𝐁𝐊C~2𝐌𝐑1𝐁𝐑]|Row {1:NO},\begin{split}\mathbf{M}_{13}&=\frac{a}{2}\left.\left[\widetilde{C}_{1}\mathbf{M_{K}}^{-1}\mathbf{B_{K}}-\widetilde{C}_{2}\mathbf{M_{R}}^{-1}\mathbf{B_{R}}\right]\right|_{\mbox{Row~{}}\{1:N_{O}\}},\\ \mathbf{M}_{23}&=-\frac{a}{3}\left.\left[\widetilde{C}_{1}\mathbf{M_{K}}^{-1}\mathbf{B_{K}}-\widetilde{C}_{2}\mathbf{M_{R}}^{-1}\mathbf{B_{R}}\right]\right|_{\mbox{Row~{}}\{1:N_{O}\}},\end{split} (47)

where the original matrices on the right hand side should be of size NA×NON_{A}\times N_{O} but only the first NON_{O} rows are utilized to form 𝐌13\mathbf{M}_{13} and 𝐌23\mathbf{M}_{23} which correspond to KA,1,,KA,NOK_{A,1},\ldots,K_{A,N_{O}} and RA,1,,RA,NOR_{A,1},\ldots,R_{A,N_{O}} within the Pacific band.

Finally, let’s consider the SST model,

tT=bT+c1η(KO+RO),\partial_{t}T=-bT+c_{1}\eta(K_{O}+R_{O}),

the discrete form of which is straightforward,

tTi=bTi+c1η(KO,i+RO,i).\partial_{t}T_{i}=-bT_{i}+c_{1}\eta(K_{O,i}+R_{O,i}).

The matrices 𝐌31\mathbf{M}_{31}, 𝐌32\mathbf{M}_{32} and 𝐌33\mathbf{M}_{33} are all diagonal and their (i,i)(i,i)-th diagonal entries are given by

(𝐌31)ii=(𝐌32)ii=c1η(xi),(𝐌33)ii=b.(\mathbf{M}_{31})_{ii}=(\mathbf{M}_{32})_{ii}=c_{1}\eta(x_{i}),\qquad(\mathbf{M}_{33})_{ii}=-b. (48)

Collecting the results in (43), (47) and (48) yields the matrix 𝐌\mathbf{M}.

7.3 Proofs of the propositions

7.3.1 Proof of Proposition 5.3

Proof.

Let a~=a+a¯\widetilde{a}=a+\bar{a} and b~=b+b¯\widetilde{b}=b+\bar{b}, where a=vxaa=vx_{a}, a¯=v¯x¯a\bar{a}=\bar{v}\bar{x}_{a}, b=vxbb=vx_{b}, and b¯=v¯x¯b\bar{b}=\bar{v}\bar{x}_{b}. It is clear that aa and bb are complex scalars. For convenience, write

a=aRe+iaIm,andb=bRe+ibIm,a=a_{Re}+ia_{Im},\qquad\mbox{and}\qquad b=b_{Re}+ib_{Im}, (49)

and therefore

a¯=aReiaIm,andb¯=bReibIma~=a+a¯=2aRe,andb~=b+b¯=2bRe.\begin{split}\bar{a}=a_{Re}-ia_{Im},&\qquad\mbox{and}\qquad\bar{b}=b_{Re}-ib_{Im}\\ \widetilde{a}=a+\bar{a}=2a_{Re},&\qquad\mbox{and}\qquad\widetilde{b}=b+\bar{b}=2b_{Re}.\end{split}

Once aRea_{Re} is computed, multiplying it by a factor of 22 finishes the solutions a~\widetilde{a} and b~\widetilde{b}. For this reason, the focus here is on solving aRea_{Re}.

Multiplying (22) by xa{x}_{a} and xb{x}_{b}, respectively, yields

dadt\displaystyle\frac{{\,\rm d}a}{{\,\rm d}t} =(do+iωo)a+ga,\displaystyle=(-d_{o}+i\omega_{o})a+g_{a}, (50a)
dbdt\displaystyle\frac{{\,\rm d}b}{{\,\rm d}t} =(do+iωo)b+gb,\displaystyle=(-d_{o}+i\omega_{o})b+g_{b}, (50b)

where ga=gxa=ga,Re+iga,Img_{a}=gx_{a}=g_{a,Re}+ig_{a,Im} and gb=gxb=gb,Re+igb,Img_{b}=gx_{b}=g_{b,Re}+ig_{b,Im}. Inserting (49) into (50) results in four equations for the real and imaginary parts of aa and the real and imaginary parts of bb,

daRedt\displaystyle\frac{{\,\rm d}a_{Re}}{{\,\rm d}t} =doaReωoaIm+ga,Re,\displaystyle=-d_{o}a_{Re}-\omega_{o}a_{Im}+g_{a,Re}, (51a)
daImdt\displaystyle\frac{{\,\rm d}a_{Im}}{{\,\rm d}t} =doaIm+ωoaRe+ga,Im,\displaystyle=-d_{o}a_{Im}+\omega_{o}a_{Re}+g_{a,Im}, (51b)
dbRedt\displaystyle\frac{{\,\rm d}b_{Re}}{{\,\rm d}t} =dobReωobIm+gb,Re,\displaystyle=-d_{o}b_{Re}-\omega_{o}b_{Im}+g_{b,Re}, (51c)
dbImdt\displaystyle\frac{{\,\rm d}b_{Im}}{{\,\rm d}t} =dobIm+ωobRe+gb,Im.\displaystyle=-d_{o}b_{Im}+\omega_{o}b_{Re}+g_{b,Im}. (51d)

Here, only the two equations (51a) and (51c) for the real part of aa and bb need to be solved, since 2aRe2a_{Re} and 2bRe2b_{Re} give the pair of physical variables we aim at.

The goal now is to use aRea_{Re} and bReb_{Re} to represent the term aIma_{Im} and bImb_{Im} in (51a) and (51c). To this end, let’s introduce two new complex scalars

c=cRe+cIm:=xaxb,andd=dRe+dIm:=xbxa,c=c_{Re}+c_{Im}:=\frac{x_{a}}{x_{b}},\qquad\mbox{and}\qquad d=d_{Re}+d_{Im}:=\frac{x_{b}}{x_{a}}, (52)

which are the ratio (and the inverse ratio) of the two bases xax_{a} and xbx_{b}. Combining (49) and (52) yields,

aRe+iaIm=(bRe+ibIm)(cRe+icIm)=(bRecRebImcIm)+i(bRecIm+bImcRe),bRe+ibIm=(aRe+iaIm)(dRe+idIm)=(aRedReaImdIm)+i(aRedIm+aImdRe),\begin{split}a_{Re}+ia_{Im}&=(b_{Re}+ib_{Im})(c_{Re}+ic_{Im})=(b_{Re}c_{Re}-b_{Im}c_{Im})+i(b_{Re}c_{Im}+b_{Im}c_{Re}),\\ b_{Re}+ib_{Im}&=(a_{Re}+ia_{Im})(d_{Re}+id_{Im})=(a_{Re}d_{Re}-a_{Im}d_{Im})+i(a_{Re}d_{Im}+a_{Im}d_{Re}),\end{split}

which leads to

aIm\displaystyle a_{Im} =cImbRe+cRebIm,\displaystyle=c_{Im}b_{Re}+c_{Re}b_{Im}, (53a)
bIm\displaystyle b_{Im} =dImaRe+dReaIm.\displaystyle=d_{Im}a_{Re}+d_{Re}a_{Im}. (53b)

With (53) in hand, plugging (53a) into (51a) yields

daRedt=doaReωocImbReωocRebIm+ga,Re.\frac{{\,\rm d}a_{Re}}{{\,\rm d}t}=-d_{o}a_{Re}-\omega_{o}c_{Im}b_{Re}-\omega_{o}c_{Re}b_{Im}+g_{a,Re}. (54)

Next, plugging (53b) into (54) to cancel bImb_{Im} leads to

daRedt=doaReωocImbReωocRedImaReωocRedReaIm+ga,Re.\frac{{\,\rm d}a_{Re}}{{\,\rm d}t}=-d_{o}a_{Re}-\omega_{o}c_{Im}b_{Re}-\omega_{o}c_{Re}d_{Im}a_{Re}-\omega_{o}c_{Re}d_{Re}a_{Im}+g_{a,Re}. (55)

Although (55) again introduces aIma_{Im}, making use of (51a) and (55) suffices to cancel this aIma_{Im}. To this end, multiplying the equation (51a) by cRedRec_{Re}d_{Re} gives

cRedRedaRedt=docRedReaReωocRedReaIm+cRedRega,Re.c_{Re}d_{Re}\frac{{\,\rm d}a_{Re}}{{\,\rm d}t}=-d_{o}c_{Re}d_{Re}a_{Re}-\omega_{o}c_{Re}d_{Re}a_{Im}+c_{Re}d_{Re}g_{a,Re}. (56)

Subtracting (56) from (55) results in

(1cRedRe)daRedt=do(1cRedRe)aReωocImbReωocRedImaRe+(1cRedRe)ga,Re,(1-c_{Re}d_{Re})\frac{{\,\rm d}a_{Re}}{{\,\rm d}t}=-d_{o}(1-c_{Re}d_{Re})a_{Re}-\omega_{o}c_{Im}b_{Re}-\omega_{o}c_{Re}d_{Im}a_{Re}+(1-c_{Re}d_{Re})g_{a,Re},

which implies

daRedt=doaReωocIm1cRedRebReωocRedIm1cRedReaRe+ga,Re.\frac{{\,\rm d}a_{Re}}{{\,\rm d}t}=-d_{o}a_{Re}-\frac{\omega_{o}c_{Im}}{1-c_{Re}d_{Re}}b_{Re}-\frac{\omega_{o}c_{Re}d_{Im}}{1-c_{Re}d_{Re}}a_{Re}+g_{a,Re}. (57)

Applying the same argument leads to the corresponding equation of bReb_{Re} that does not explicitly depend on bImb_{Im},

dbRedt=dobReωodIm1dRecReaReωodRecIm1dRecRebRe+gb,Re.\frac{{\,\rm d}b_{Re}}{{\,\rm d}t}=-d_{o}b_{Re}-\frac{\omega_{o}d_{Im}}{1-d_{Re}c_{Re}}a_{Re}-\frac{\omega_{o}d_{Re}c_{Im}}{1-d_{Re}c_{Re}}b_{Re}+g_{b,Re}. (58)

Finally, multiplying both aRea_{Re} and bReb_{Re} by a factor of 22 gives the two-dimensional oscillator model of a~\tilde{a} and b~\tilde{b},

da~dt=doa~ωocIm1cRedReb~ωocRedIm1cRedRea~+2ga,Re,db~dt=dob~ωodIm1dRecRea~ωodRecIm1dRecReb~+2gb,Re,\begin{split}\frac{{\,\rm d}\widetilde{a}}{{\,\rm d}t}&=-d_{o}\widetilde{a}-\frac{\omega_{o}c_{Im}}{1-c_{Re}d_{Re}}\widetilde{b}-\frac{\omega_{o}c_{Re}d_{Im}}{1-c_{Re}d_{Re}}\widetilde{a}+2g_{a,Re},\\ \frac{{\,\rm d}\widetilde{b}}{{\,\rm d}t}&=-d_{o}\widetilde{b}-\frac{\omega_{o}d_{Im}}{1-d_{Re}c_{Re}}\widetilde{a}-\frac{\omega_{o}d_{Re}c_{Im}}{1-d_{Re}c_{Re}}\widetilde{b}+2g_{b,Re},\end{split} (59)

where

ga,Re=Re(gxa)=Re(𝐳T𝐬𝐮xa)apandgb,Re=Re(gxb)=Re(𝐳T𝐬𝐮xb)ap.g_{a,Re}=Re(gx_{a})=Re(\mathbf{z}^{T}\mathbf{s}_{\mathbf{u}}x_{a})a_{p}\qquad\mbox{and}\qquad g_{b,Re}=Re(gx_{b})=Re(\mathbf{z}^{T}\mathbf{s}_{\mathbf{u}}x_{b})a_{p}. (60)

It is convenient to define

C~=(c~11c~12c~21c~22)=ωo1cRedRe(cRedImcImdImdRecIm)=ωo1cos2(αβ)(sin(αβ)cos(αβ)AaAbsin(αβ)AbAasin(αβ)sin(αβ)cos(αβ)),\begin{split}\widetilde{C}&=\left(\begin{array}[]{cc}\widetilde{c}_{11}&\widetilde{c}_{12}\\ \widetilde{c}_{21}&\widetilde{c}_{22}\\ \end{array}\right)=-\frac{\omega_{o}}{1-c_{Re}d_{Re}}\left(\begin{array}[]{cc}c_{Re}d_{Im}&c_{Im}\\ d_{Im}&d_{Re}c_{Im}\\ \end{array}\right)\\ &=-\frac{\omega_{o}}{1-\cos^{2}(\alpha-\beta)}\left(\begin{array}[]{cc}-\sin(\alpha-\beta)\cos(\alpha-\beta)&\frac{A_{a}}{A_{b}}\sin(\alpha-\beta)\\ -\frac{A_{b}}{A_{a}}\sin(\alpha-\beta)&\sin(\alpha-\beta)\cos(\alpha-\beta)\\ \end{array}\right),\end{split} (61)

which links (c,d)(c,d) with (α,β)(\alpha,\beta) and gives the form in (29). ∎

7.3.2 Proof of Proposition 5.4

Proof.

The delayed oscillator is formed based on the same starting idea from the discharge-recharge oscillator. It exploits the equations in (59) to represent b~(τ)\tilde{b}(\tau) in the first equation as a function of a~(τδ)\tilde{a}(\tau-\delta) leads to a delayed El Niño.

Let us start with (59). To make the notation simple, rewrite (59) as follows:

da~dt=doa~+c~11a~+c~12b~+g~a,db~dt=dob~+c~21a~+c~22b~+g~b,\begin{split}\frac{{\,\rm d}\widetilde{a}}{{\,\rm d}t}&=-d_{o}\widetilde{a}+\widetilde{c}_{11}\widetilde{a}+\widetilde{c}_{12}\widetilde{b}+\widetilde{g}_{a},\\ \frac{{\,\rm d}\widetilde{b}}{{\,\rm d}t}&=-d_{o}\widetilde{b}+\widetilde{c}_{21}\widetilde{a}+\widetilde{c}_{22}\widetilde{b}+\widetilde{g}_{b},\end{split} (62)

where g~a=2ga,Re\widetilde{g}_{a}=2g_{a,Re} and g~b=2gb,Re\widetilde{g}_{b}=2g_{b,Re}. Define two-dimensional vectors u~,g~\widetilde{u},\widetilde{g} and a 2×22\times 2 matrix 𝐂{\mathbf{C}},

𝐮=(a~b~),𝐠=(g~ag~b),𝐂=do𝐈+(c~11c~12c~21c~22)=do𝐈+C~.{\mathbf{u}}=\left(\begin{array}[]{c}\widetilde{a}\\ \widetilde{b}\\ \end{array}\right),\qquad{\mathbf{g}}=\left(\begin{array}[]{c}\widetilde{g}_{a}\\ \widetilde{g}_{b}\\ \end{array}\right),\qquad{\mathbf{C}}=-d_{o}\mathbf{I}+\left(\begin{array}[]{cc}\widetilde{c}_{11}&\widetilde{c}_{12}\\ \widetilde{c}_{21}&\widetilde{c}_{22}\\ \end{array}\right)=-d_{o}\mathbf{I}+\widetilde{C}.

It is easy to check that 𝐂{\mathbf{C}} has two distinct eigenvectors do±iωo-d_{o}\pm i\omega_{o}. Thus, a similarity transform leads to 𝐂=𝐘𝚲𝐘1{\mathbf{C}}=\mathbf{Y}{\mathbf{\Lambda}}\mathbf{Y}^{-1}, where 𝚲{\mathbf{\Lambda}} is a diagonal matrix with diagonal entries being do±iωo-d_{o}\pm i\omega_{o}. Now, (62) can be written as

d𝐮dt=𝐂𝐮+𝐠=𝐘𝚲𝐘1𝐮+𝐠.\frac{{\,\rm d}{\mathbf{u}}}{{\,\rm d}t}={\mathbf{C}}{\mathbf{u}}+{\mathbf{g}}=\mathbf{Y}{\mathbf{\Lambda}}\mathbf{Y}^{-1}{\mathbf{u}}+{\mathbf{g}}. (63)

Define 𝐰=𝐘1𝐮{\mathbf{w}}=\mathbf{Y}^{-1}{\mathbf{u}} and 𝐟=𝐘1𝐠{\mathbf{f}}=\mathbf{Y}^{-1}{\mathbf{g}}. Multiplying (63) by 𝐘1\mathbf{Y}^{-1} yields

d𝐰dt=𝚲𝐰+𝐟,\frac{{\,\rm d}{\mathbf{w}}}{{\,\rm d}t}={\mathbf{\Lambda}}{\mathbf{w}}+{\mathbf{f}},

where

𝐰=(ww¯),𝐟=(ff¯),𝚲=(do+iωo00doiωo),{\mathbf{w}}=\left(\begin{array}[]{c}w\\ \bar{w}\\ \end{array}\right),\qquad{\mathbf{f}}=\left(\begin{array}[]{c}f\\ \bar{f}\\ \end{array}\right),\qquad{\mathbf{\Lambda}}=\left(\begin{array}[]{cc}-d_{o}+i\omega_{o}&0\\ 0&-d_{o}-i\omega_{o}\\ \end{array}\right),

and therefore

dwdt=(do+iωo)w+f.\frac{{\,\rm d}w}{{\,\rm d}t}=(-d_{o}+i\omega_{o})w+f. (64)

Once 𝐰\mathbf{w} is solved, 𝐮\mathbf{u} is easily recovered by 𝐮=𝐘𝐰\mathbf{u}=\mathbf{Y}\mathbf{w}. Since the two eigenvalues are a complex conjugate pair, the eigenvector matrix 𝐘\mathbf{Y} must have the following form

𝐘=(y1y¯1y2y¯2).\mathbf{Y}=\left(\begin{array}[]{cc}y_{1}&\bar{y}_{1}\\ y_{2}&\bar{y}_{2}\\ \end{array}\right). (65)

Therefore,

a~:=u1=2Re(y1w),andb~:=u2=2Re(y2w).\widetilde{a}:=u_{1}=2Re(y_{1}w),\qquad\mbox{and}\qquad\widetilde{b}:=u_{2}=2Re(y_{2}w). (66)

For the notation simplicity below, rewrite y1y_{1} and y2y_{2} into the following form

y1=y1,Re+iy1,Im,andy2=y2,Re+iy2,Im.y_{1}=y_{1,Re}+iy_{1,Im},\qquad\mbox{and}\qquad y_{2}=y_{2,Re}+iy_{2,Im}.

Similarly, rewrite the initial value w(0)w(0) and the forcing f(s)f(s) as

w(0)=wRe(0)+iwIm(0),andf(s)=fRe(0)+ifIm(0).w(0)=w_{Re}(0)+iw_{Im}(0),\qquad\mbox{and}\qquad f(s)=f_{Re}(0)+if_{Im}(0).

Thus, the solution of (64) can be written as

w(t)=w(0)edot(cos(ωot)+isin(ωot))+0tedo(ts)(cos(ωo(ts))+isin(ωo(ts)))f(s)𝑑s=edot[(wRe(0)cos(ωot)wIm(0)sin(ωot))+i(wRe(0)sin(ωot)+wIm(0)cos(ωot))]+0tedo(ts)[(cos(ωo(ts))fRe(s)sin(ωo(ts))fIm)+i(cos(ωo(ts))fIm(s)+sin(ωo(ts))fRe)]ds.\begin{split}w(t)&=w(0)e^{-d_{o}t}\Big{(}\cos(\omega_{o}t)+i\sin(\omega_{o}t)\Big{)}\\ &\qquad\qquad+\int_{0}^{t}e^{-d_{o}(t-s)}\Big{(}\cos\big{(}\omega_{o}(t-s)\big{)}+i\sin\big{(}\omega_{o}(t-s)\big{)}\Big{)}f(s)ds\\ &=e^{-d_{o}t}\Big{[}\Big{(}w_{Re}(0)\cos(\omega_{o}t)-w_{Im}(0)\sin(\omega_{o}t)\Big{)}\\ &\qquad\qquad\qquad\qquad\qquad+i\Big{(}w_{Re}(0)\sin(\omega_{o}t)+w_{Im}(0)\cos(\omega_{o}t)\Big{)}\Big{]}\\ &\qquad+\int_{0}^{t}e^{-d_{o}(t-s)}\Big{[}\Big{(}\cos\big{(}\omega_{o}(t-s)\big{)}f_{Re}(s)-\sin\big{(}\omega_{o}(t-s)\big{)}f_{Im}\Big{)}\\ &\qquad\qquad+i\Big{(}\cos\big{(}\omega_{o}(t-s)\big{)}f_{Im}(s)+\sin\big{(}\omega_{o}(t-s)\big{)}f_{Re}\Big{)}\Big{]}ds.\end{split}

Therefore,

u1(t)=2Re(y1w)=2edot[y1,Re(wRe(0)cos(ωot)wIm(0)sin(ωot))y1,Im(wRe(0)sin(ωot)+wIm(0)cos(ωot))]+20tedo(ts)[y1,Re(cos(ωo(ts))fRe(s)sin(ωo(ts))fIm(s))y1,Im(cos(ωo(ts))fIm(s)+sin(ωo(ts))fRe(s))]ds.\begin{split}u_{1}(t)&=2Re(y_{1}w)\\ &=2e^{-d_{o}t}\Big{[}y_{1,Re}\Big{(}w_{Re}(0)\cos(\omega_{o}t)-w_{Im}(0)\sin(\omega_{o}t)\Big{)}\\ &\qquad\qquad\qquad\qquad\qquad\qquad-y_{1,Im}\Big{(}w_{Re}(0)\sin(\omega_{o}t)+w_{Im}(0)\cos(\omega_{o}t)\Big{)}\Big{]}\\ &\qquad+2\int_{0}^{t}e^{-d_{o}(t-s)}\Big{[}y_{1,Re}\Big{(}\cos\big{(}\omega_{o}(t-s)\big{)}f_{Re}(s)-\sin\big{(}\omega_{o}(t-s)\big{)}f_{Im}(s)\Big{)}\\ &\qquad\qquad-y_{1,Im}\Big{(}\cos\big{(}\omega_{o}(t-s)\big{)}f_{Im}(s)+\sin\big{(}\omega_{o}(t-s)\big{)}f_{Re}(s)\Big{)}\Big{]}ds.\end{split} (67)

Similarly,

u2(t)=2Re(y2w)=2edot[y2,Re(wRe(0)cos(ωot)wIm(0)sin(ωot))y2,Im(wRe(0)sin(ωot)+wIm(0)cos(ωot))]+20tedo(ts)[y2,Re(cos(ωo(ts))fRe(s)sin(ωo(ts))fIm(s))y2,Im(cos(ωo(ts))fIm(s)+sin(ωo(ts))fRe(s))]ds.\begin{split}u_{2}(t)&=2Re(y_{2}w)\\ &=2e^{-d_{o}t}\Big{[}y_{2,Re}\Big{(}w_{Re}(0)\cos(\omega_{o}t)-w_{Im}(0)\sin(\omega_{o}t)\Big{)}\\ &\qquad\qquad\qquad\qquad\qquad\qquad-y_{2,Im}\Big{(}w_{Re}(0)\sin(\omega_{o}t)+w_{Im}(0)\cos(\omega_{o}t)\Big{)}\Big{]}\\ &\qquad+2\int_{0}^{t}e^{-d_{o}(t-s)}\Big{[}y_{2,Re}\Big{(}\cos\big{(}\omega_{o}(t-s)\big{)}f_{Re}(s)-\sin\big{(}\omega_{o}(t-s)\big{)}f_{Im}(s)\Big{)}\\ &\qquad\qquad-y_{2,Im}\Big{(}\cos\big{(}\omega_{o}(t-s)\big{)}f_{Im}(s)+\sin\big{(}\omega_{o}(t-s)\big{)}f_{Re}(s)\Big{)}\Big{]}ds.\end{split} (68)

Note that (67) and (68) have the same composition: a contribution from initial value and a contribution from external forcing. Below, let’s write down these two components (and drop the common prefactors) for the convenience of discussion.

I. Composition of initial value:

u1I=(y1,RewRe(0)y1,ImwIm(0))cos(ωot)(y1,RewIm(0)y1,ImwRe(0))sin(ωot)u2I=(y2,RewRe(0)y2,ImwIm(0))cos(ωot)(y2,RewIm(0)y2,ImwRe(0))sin(ωot)\begin{split}u_{1}^{I}=&\Big{(}y_{1,Re}w_{Re}(0)-y_{1,Im}w_{Im}(0)\Big{)}\cos(\omega_{o}t)-\Big{(}y_{1,Re}w_{Im}(0)-y_{1,Im}w_{Re}(0)\Big{)}\sin(\omega_{o}t)\\ u_{2}^{I}=&\Big{(}y_{2,Re}w_{Re}(0)-y_{2,Im}w_{Im}(0)\Big{)}\cos(\omega_{o}t)-\Big{(}y_{2,Re}w_{Im}(0)-y_{2,Im}w_{Re}(0)\Big{)}\sin(\omega_{o}t)\end{split}

II. Composition of external forcing:

u1II=(y1,RefRe(s)y1,ImfIm(s))cos(ωo(ts))(y1,RefIm(s)y1,ImfRe(s))sin(ωo(ts))u2II=(y2,RefRe(s)y2,ImfIm(s))cos(ωo(ts))(y2,RefIm(s)y2,ImfRe(s))sin(ωo(ts))\begin{split}u_{1}^{II}=&\Big{(}y_{1,Re}f_{Re}(s)-y_{1,Im}f_{Im}(s)\Big{)}\cos(\omega_{o}(t-s))-\Big{(}y_{1,Re}f_{Im}(s)-y_{1,Im}f_{Re}(s)\Big{)}\sin(\omega_{o}(t-s))\\ u_{2}^{II}=&\Big{(}y_{2,Re}f_{Re}(s)-y_{2,Im}f_{Im}(s)\Big{)}\cos(\omega_{o}(t-s))-\Big{(}y_{2,Re}f_{Im}(s)-y_{2,Im}f_{Re}(s)\Big{)}\sin(\omega_{o}(t-s))\end{split}

For the initial value part, rewrite the two expressions as

u1I\displaystyle u_{1}^{I} =(y1,RewRe(0)y1,ImwIm(0)y1,RewIm(0)y1,ImwRe(0))(cos(ωot)sin(ωot))\displaystyle=\left(\begin{array}[]{cc}y_{1,Re}w_{Re}(0)-y_{1,Im}w_{Im}(0)&-y_{1,Re}w_{Im}(0)-y_{1,Im}w_{Re}(0)\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}t)\\ \sin(\omega_{o}t)\\ \end{array}\right) (69d)
=(wRe(0)wIm(0))(y1,Rey1,Imy1,Imy1,Re)(cos(ωot)sin(ωot))\displaystyle=\left(\begin{array}[]{cc}w_{Re}(0)&w_{Im}(0)\\ \end{array}\right)\left(\begin{array}[]{cc}y_{1,Re}&-y_{1,Im}\\ -y_{1,Im}&-y_{1,Re}\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}t)\\ \sin(\omega_{o}t)\\ \end{array}\right) (69j)
u2I\displaystyle u_{2}^{I} =(y2,RewRe(0)y2,ImwIm(0)y2,RewIm(0)y2,ImwRe(0))(cos(ωot)sin(ωot))\displaystyle=\left(\begin{array}[]{cc}y_{2,Re}w_{Re}(0)-y_{2,Im}w_{Im}(0)&-y_{2,Re}w_{Im}(0)-y_{2,Im}w_{Re}(0)\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}t)\\ \sin(\omega_{o}t)\\ \end{array}\right) (69n)
=(wRe(0)wIm(0))(y2,Rey2,Imy2,Imy2,Re)(cos(ωot)sin(ωot))\displaystyle=\left(\begin{array}[]{cc}w_{Re}(0)&w_{Im}(0)\\ \end{array}\right)\left(\begin{array}[]{cc}y_{2,Re}&-y_{2,Im}\\ -y_{2,Im}&-y_{2,Re}\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}t)\\ \sin(\omega_{o}t)\\ \end{array}\right) (69t)

and

u1II\displaystyle u_{1}^{II} =(y1,RefRe(s)y1,ImfIm(s)y1,RefIm(s)y1,ImfRe(s))(cos(ωo(ts))sin(ωo(ts)))\displaystyle=\left(\begin{array}[]{cc}y_{1,Re}f_{Re}(s)-y_{1,Im}f_{Im}(s)&-y_{1,Re}f_{Im}(s)-y_{1,Im}f_{Re}(s)\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}(t-s))\\ \sin(\omega_{o}(t-s))\\ \end{array}\right) (70d)
=(fRe(s)fIm(s))(y1,Rey1,Imy1,Imy1,Re)(cos(ωo(ts))sin(ωo(ts)))\displaystyle=\left(\begin{array}[]{cc}f_{Re}(s)&f_{Im}(s)\\ \end{array}\right)\left(\begin{array}[]{cc}y_{1,Re}&-y_{1,Im}\\ -y_{1,Im}&-y_{1,Re}\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}(t-s))\\ \sin(\omega_{o}(t-s))\\ \end{array}\right) (70j)
u2II\displaystyle u_{2}^{II} =(y2,RefRe(s)y2,ImfIm(s)y2,RefIm(s)y2,ImfRe(s))(cos(ωo(ts))sin(ωo(ts)))\displaystyle=\left(\begin{array}[]{cc}y_{2,Re}f_{Re}(s)-y_{2,Im}f_{Im}(s)&-y_{2,Re}f_{Im}(s)-y_{2,Im}f_{Re}(s)\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}(t-s))\\ \sin(\omega_{o}(t-s))\\ \end{array}\right) (70n)
=(fRe(s)fIm(s))(y2,Rey2,Imy2,Imy2,Re)(cos(ωo(ts))sin(ωo(ts)))\displaystyle=\left(\begin{array}[]{cc}f_{Re}(s)&f_{Im}(s)\\ \end{array}\right)\left(\begin{array}[]{cc}y_{2,Re}&-y_{2,Im}\\ -y_{2,Im}&-y_{2,Re}\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}(t-s))\\ \sin(\omega_{o}(t-s))\\ \end{array}\right) (70t)

It is clear that the difference between u1u_{1} and u2u_{2} is given by the angle φ\varphi between the two matrices consisted of the eigenvectors y1,y2y_{1},y_{2}

My1=(y1,Rey1,Imy1,Imy1,Re)andMy2=(y2,Rey2,Imy2,Imy2,Re),M_{y_{1}}=\left(\begin{array}[]{cc}y_{1,Re}&-y_{1,Im}\\ -y_{1,Im}&-y_{1,Re}\\ \end{array}\right)\qquad\mbox{and}\qquad M_{y_{2}}=\left(\begin{array}[]{cc}y_{2,Re}&-y_{2,Im}\\ -y_{2,Im}&-y_{2,Re}\\ \end{array}\right), (71)

that is, there exists a rotation matrix and a constant

Φ=(cosφsinφsinφcosφ),cy=y2y1\varPhi=\left(\begin{array}[]{cc}\cos\varphi&\sin\varphi\\ -\sin\varphi&\cos\varphi\\ \end{array}\right),\qquad c_{y}=\frac{\|y_{2}\|}{\|y_{1}\|} (72)

such that

My2=cyMy1Φ.M_{y_{2}}=c_{y}M_{y_{1}}\varPhi. (73)

Making use of (73), the formula in (69t) becomes

u2I=(wRe(0)wIm(0))(y2,Rey2,Imy2,Imy2,Re)(cos(ωot)sin(ωot))=cy(wRe(0)wIm(0))(y1,Rey1,Imy1,Imy1,Re)(cosφsinφsinφcosφ)(cos(ωot)sin(ωot))=cy(wRe(0)wIm(0))(y1,Rey1,Imy1,Imy1,Re)(cosφcos(ωot)+sinφsin(ωot)sinφcos(ωot)+cosφsin(ωot))=cy(wRe(0)wIm(0))(y1,Rey1,Imy1,Imy1,Re)(cos(ωotφ)sin(ωotφ))\begin{split}u_{2}^{I}&=\left(\begin{array}[]{cc}w_{Re}(0)&w_{Im}(0)\\ \end{array}\right)\left(\begin{array}[]{cc}y_{2,Re}&-y_{2,Im}\\ -y_{2,Im}&-y_{2,Re}\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}t)\\ \sin(\omega_{o}t)\\ \end{array}\right)\\ &=c_{y}\left(\begin{array}[]{cc}w_{Re}(0)&w_{Im}(0)\\ \end{array}\right)\left(\begin{array}[]{cc}y_{1,Re}&-y_{1,Im}\\ -y_{1,Im}&-y_{1,Re}\\ \end{array}\right)\left(\begin{array}[]{cc}\cos\varphi&\sin\varphi\\ -\sin\varphi&\cos\varphi\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}t)\\ \sin(\omega_{o}t)\\ \end{array}\right)\\ &=c_{y}\left(\begin{array}[]{cc}w_{Re}(0)&w_{Im}(0)\\ \end{array}\right)\left(\begin{array}[]{cc}y_{1,Re}&-y_{1,Im}\\ -y_{1,Im}&-y_{1,Re}\\ \end{array}\right)\left(\begin{array}[]{c}\cos\varphi\cos(\omega_{o}t)+\sin\varphi\sin(\omega_{o}t)\\ -\sin\varphi\cos(\omega_{o}t)+\cos\varphi\sin(\omega_{o}t)\\ \end{array}\right)\\ &=c_{y}\left(\begin{array}[]{cc}w_{Re}(0)&w_{Im}(0)\\ \end{array}\right)\left(\begin{array}[]{cc}y_{1,Re}&-y_{1,Im}\\ -y_{1,Im}&-y_{1,Re}\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}t-\varphi)\\ \sin(\omega_{o}t-\varphi)\\ \end{array}\right)\end{split} (74)

Similarly, (70t) can be rewritten as

u2II=(fRe(s)fIm(s))(y2,Rey2,Imy2,Imy2,Re)(cos(ωo(ts))sin(ωo(ts)))=cy(fRe(s)fIm(s))(y1,Rey1,Imy1,Imy1,Re)(cos(ωo(ts)φ)sin(ωo(ts)φ))\begin{split}u_{2}^{II}&=\left(\begin{array}[]{cc}f_{Re}(s)&f_{Im}(s)\\ \end{array}\right)\left(\begin{array}[]{cc}y_{2,Re}&-y_{2,Im}\\ -y_{2,Im}&-y_{2,Re}\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}(t-s))\\ \sin(\omega_{o}(t-s))\\ \end{array}\right)\\ &=c_{y}\left(\begin{array}[]{cc}f_{Re}(s)&f_{Im}(s)\\ \end{array}\right)\left(\begin{array}[]{cc}y_{1,Re}&-y_{1,Im}\\ -y_{1,Im}&-y_{1,Re}\\ \end{array}\right)\left(\begin{array}[]{c}\cos(\omega_{o}(t-s)-\varphi)\\ \sin(\omega_{o}(t-s)-\varphi)\\ \end{array}\right)\end{split} (75)

Therefore, in light of (74) and (75), (68) becomes

u2(t)=2edot[y2,Re(wRe(0)cos(ωot)wIm(0)sin(ωot))y2,Im(wRe(0)sin(ωot)+wIm(0)cos(ωot))]+20tedo(ts)[y2,Re(cos(ωo(ts))fRe(s)sin(ωo(ts))fIm(s))y2,Im(cos(ωo(ts))fIm(s)+sin(ωo(ts))fRe(s))]ds=2cyedot[y1,Re(wRe(0)cos(ωotφ)wIm(0)sin(ωotφ))y1,Im(wRe(0)sin(ωot)+wIm(0)cos(ωot))]+2cy0tedo(ts)[y1,Re(cos(ωo(ts)φ)fRe(s)sin(ωo(ts)φ)fIm(s))y1,Im(cos(ωo(ts)φ)fIm(s)+sin(ωo(ts)φ)fRe(s))]ds.\begin{split}u_{2}(t)&=2e^{-d_{o}t}\Big{[}y_{2,Re}\Big{(}w_{Re}(0)\cos(\omega_{o}t)-w_{Im}(0)\sin(\omega_{o}t)\Big{)}\\ &\qquad\qquad\qquad\qquad\qquad\qquad-y_{2,Im}\Big{(}w_{Re}(0)\sin(\omega_{o}t)+w_{Im}(0)\cos(\omega_{o}t)\Big{)}\Big{]}\\ &\qquad+2\int_{0}^{t}e^{-d_{o}(t-s)}\Big{[}y_{2,Re}\Big{(}\cos\big{(}\omega_{o}(t-s)\big{)}f_{Re}(s)-\sin\big{(}\omega_{o}(t-s)\big{)}f_{Im}(s)\Big{)}\\ &\qquad\qquad-y_{2,Im}\Big{(}\cos\big{(}\omega_{o}(t-s)\big{)}f_{Im}(s)+\sin\big{(}\omega_{o}(t-s)\big{)}f_{Re}(s)\Big{)}\Big{]}ds\\ &=2c_{y}e^{-d_{o}t}\Big{[}y_{1,Re}\Big{(}w_{Re}(0)\cos(\omega_{o}t-\varphi)-w_{Im}(0)\sin(\omega_{o}t-\varphi)\Big{)}\\ &\qquad\qquad\qquad-y_{1,Im}\Big{(}w_{Re}(0)\sin(\omega_{o}t)+w_{Im}(0)\cos(\omega_{o}t)\Big{)}\Big{]}\\ &\quad+2c_{y}\int_{0}^{t}e^{-d_{o}(t-s)}\Big{[}y_{1,Re}\Big{(}\cos\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Re}(s)-\sin\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Im}(s)\Big{)}\\ &\qquad\qquad-y_{1,Im}\Big{(}\cos\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Im}(s)+\sin\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Re}(s)\Big{)}\Big{]}ds.\end{split} (76)

Comparing (67) and (76) leads to

u2(t)=cyedoφ/ωou1(tφ/ωo)+2cytφ/ωotedo(ts)[y1,Re(cos(ωo(ts)φ)fRe(s)sin(ωo(ts)φ)fIm(s))y1,Im(cos(ωo(ts)φ)fIm(s)+sin(ωo(ts)φ)fRe(s))]ds\begin{split}u_{2}(t)&=c_{y}e^{-d_{o}\varphi/\omega_{o}}u_{1}(t-\varphi/\omega_{o})\\ &\quad+2c_{y}\int_{t-\varphi/\omega_{o}}^{t}e^{-d_{o}(t-s)}\bigg{[}y_{1,Re}\Big{(}\cos\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Re}(s)\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\sin\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Im}(s)\Big{)}\\ &\qquad-y_{1,Im}\Big{(}\cos\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Im}(s)+\sin\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Re}(s)\Big{)}\bigg{]}ds\end{split} (77)

Note that in (66) we defined a~:=u1\widetilde{a}:=u_{1} and b~:=u2\widetilde{b}:=u_{2}. Thus, plugging (77) to the first equation of (62) yields

da~(t)dt=doa~(t)+c~11a~(t)+c~12b~(t)+g~a(t)=doa~(t)+c~11a~(t)+c~12cyedoφ/ωoa~(tφ/ωo)+g~a(t)+2cyc~abtφ/ωotedo(ts)[y1,Re(cos(ωo(ts)φ)fRe(s)sin(ωo(ts)φ)fIm(s))y1,Im(cos(ωo(ts)φ)fIm(s)+sin(ωo(ts)φ)fRe(s))]ds\begin{split}\frac{{\,\rm d}\widetilde{a}(t)}{{\,\rm d}t}&=-d_{o}\widetilde{a}(t)+\widetilde{c}_{11}\widetilde{a}(t)+\widetilde{c}_{12}\widetilde{b}(t)+\widetilde{g}_{a}(t)\\ &=-d_{o}\widetilde{a}(t)+\widetilde{c}_{11}\widetilde{a}(t)+\widetilde{c}_{12}c_{y}e^{-d_{o}\varphi/\omega_{o}}\widetilde{a}(t-\varphi/\omega_{o})+\widetilde{g}_{a}(t)\\ &\quad+2c_{y}\widetilde{c}_{ab}\int_{t-\varphi/\omega_{o}}^{t}e^{-d_{o}(t-s)}\bigg{[}y_{1,Re}\Big{(}\cos\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Re}(s)\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\sin\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Im}(s)\Big{)}\\ &\qquad-y_{1,Im}\Big{(}\cos\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Im}(s)+\sin\big{(}\omega_{o}(t-s)-\varphi\big{)}f_{Re}(s)\Big{)}\bigg{]}{\,\rm d}s\end{split} (78)

References

  • [1] S. G. H. Philander, El Niño Southern Oscillation phenomena, Nature 302 (5906) (1983) 295–301.
  • [2] C. F. Ropelewski, M. S. Halpert, Global and regional scale precipitation patterns associated with the El Niño/Southern Oscillation, Monthly Weather Review 115 (8) (1987) 1606–1626.
  • [3] S. A. Klein, B. J. Soden, N.-C. Lau, Remote sea surface temperature variations during ENSO: Evidence for a tropical atmospheric bridge, Journal of Climate 12 (4) (1999) 917–932.
  • [4] M. J. McPhaden, S. E. Zebiak, M. H. Glantz, ENSO as an integrating concept in earth science, Science 314 (5806) (2006) 1740–1745.
  • [5] P. S. Schopf, M. J. Suarez, Vacillations in a coupled ocean–atmosphere model, Journal of Atmospheric Sciences 45 (3) (1988) 549–566.
  • [6] J. Picaut, M. Ioualalen, C. Menkès, T. Delcroix, M. J. Mcphaden, Mechanism of the zonal displacements of the pacific warm pool: Implications for ENSO, Science 274 (5292) (1996) 1486–1489.
  • [7] F.-F. Jin, An equatorial ocean recharge paradigm for ENSO. Part I: Conceptual model, Journal of the Atmospheric Sciences 54 (7) (1997) 811–829.
  • [8] C. Wang, J. Picaut, Understanding ENSO physics—A review, Earth’s Climate: The Ocean–Atmosphere Interaction, Geophys. Monogr 147 (2004) 21–48.
  • [9] N. Chen, X. Fang, J.-Y. Yu, A multiscale model for El Niño complexity, npj Climate and Atmospheric Science 5 (1) (2022) 1–13.
  • [10] A. Capotondi, P. D. Sardeshmukh, L. Ricciardulli, The nature of the stochastic wind forcing of ENSO, Journal of Climate 31 (19) (2018) 8081–8099.
  • [11] P. Leeuwen, Balanced ocean-data assimilation near the equator, Journal of Physical Oceanography 32 (2002) 25092519Caniaux.
  • [12] X. Fang, F. Zheng, Simulating eastern-and central-pacific type ENSO using a simple coupled model, Advances in Atmospheric Sciences 35 (6) (2018) 671–681.
  • [13] S. Thual, A. J. Majda, N. Chen, S. N. Stechmann, Simple stochastic model for El Niño with westerly wind bursts, Proceedings of the National Academy of Sciences 113 (37) (2016) 10245–10250.
  • [14] N. Chen, A. J. Majda, Simple stochastic dynamical models capturing the statistical diversity of El Niño Southern Oscillation, Proceedings of the National Academy of Sciences 114 (7) (2017) 1468–1473.
  • [15] S. E. Zebiak, M. A. Cane, A model El Niño–Southern Oscillation, Monthly Weather Review 115 (10) (1987) 2262–2278.
  • [16] D. S. Battisti, A. C. Hirst, Interannual variability in a tropical atmosphere–ocean model: Influence of the basic state, ocean geometry and nonlinearity, Journal of the Atmospheric Sciences 46 (12) (1989) 1687–1712.
  • [17] J. D. Neelin, F.-F. Jin, Modes of interannual tropical ocean–atmosphere interaction—a unified view. Part II: Analytical results in the weak-coupling limit, Journal of Atmospheric Sciences 50 (21) (1993) 3504–3522.
  • [18] R.-H. Zhang, S. E. Zebiak, R. Kleeman, N. Keenlyside, A new intermediate coupled model for El Niño simulation and prediction, Geophysical Research Letters 30 (19).
  • [19] N. Chen, X. Fang, A simple multiscale intermediate coupled stochastic model for El Niño diversity and complexity, arXiv preprint arXiv:2206.06649.
  • [20] Y. Y. Planton, E. Guilyardi, A. T. Wittenberg, J. Lee, P. J. Gleckler, T. Bayr, S. McGregor, M. J. McPhaden, S. Power, R. Roehrig, et al., Evaluating climate models with the clivar 2020 ENSO metrics package, Bulletin of the American Meteorological Society 102 (2) (2021) E193–E217.
  • [21] N.-C. Lau, M. J. Nath, Impact of ENSO on the variability of the Asian–Australian monsoons as simulated in GCM experiments, Journal of Climate 13 (24) (2000) 4287–4309.
  • [22] E. Guilyardi, A. Capotondi, M. Lengaigne, S. Thual, A. T. Wittenberg, Enso modeling: History, progress, and challenges, El Niño Southern Oscillation in a changing climate (2020) 199–226.
  • [23] B. Zhao, A. Fedorov, The effects of background zonal and meridional winds on ENSO in a coupled GCM, Journal of Climate 33 (6) (2020) 2075–2091.
  • [24] F.-F. Jin, An equatorial ocean recharge paradigm for ENSO. Part II: A stripped-down coupled model, Journal of the Atmospheric Sciences 54 (7) (1997) 830–847.
  • [25] M. Latif, D. Anderson, T. Barnett, M. Cane, R. Kleeman, A. Leetmaa, J. O’Brien, A. Rosati, E. Schneider, A review of the predictability and prediction of ENSO, Journal of Geophysical Research: Oceans 103 (C7) (1998) 14375–14393.
  • [26] Y. Xue, M. Cane, S. Zebiak, M. Blumenthal, On the prediction of ENSO: A study with a low-order Markov model, Tellus A 46 (4) (1994) 512–528.
  • [27] X. Fang, N. Chen, Quantifying the predictability of ENSO complexity using a statistically accurate multiscale stochastic model and information theory, arXiv preprint arXiv:2203.02657.
  • [28] K. Wyrtki, El Niño–the dynamic response of the equatorial Pacific ocean to atmospheric forcing, Journal of Physical Oceanography 5 (4) (1975) 572–584.
  • [29] M. J. Suarez, P. S. Schopf, A delayed action oscillator for ENSO, Journal of Atmospheric Sciences 45 (21) (1988) 3283–3287.
  • [30] J. P. McCreary Jr, A model of tropical ocean-atmosphere interaction, Monthly Weather Review 111 (2) (1983) 370–387.
  • [31] R. H. Weisberg, C. Wang, A western Pacific oscillator paradigm for the El Niño-Southern Oscillation, Geophysical Research Letters 24 (7) (1997) 779–782.
  • [32] J. Picaut, F. Masia, Y. Du Penhoat, An advective-reflective conceptual model for the oscillatory nature of the ENSO, Science 277 (5326) (1997) 663–666.
  • [33] C. Wang, A unified oscillator model for the el niño–southern oscillation, Journal of Climate 14 (1) (2001) 98–115.
  • [34] R. W. Reynolds, T. M. Smith, C. Liu, D. B. Chelton, K. S. Casey, M. G. Schlax, Daily high-resolution-blended analyses for sea surface temperature, Journal of Climate 20 (22) (2007) 5473–5496.
  • [35] D. Behringer, Y. Xue, Evaluation of the global ocean data assimilation system at NCEP: The Pacific Ocean, in: Proc. Eighth Symp. on Integrated Observing and Assimilation Systems for Atmosphere, Oceans, and Land Surface, 2004.
  • [36] E. Kalnay, M. Kanamitsu, R. Kistler, W. Collins, D. Deaven, L. Gandin, M. Iredell, S. Saha, G. White, J. Woollen, et al., The NCEP/NCAR 40-year reanalysis project, Bulletin of the American Meteorological Society 77 (3) (1996) 437–472.
  • [37] T. Matsuno, Quasi-geostrophic motions in the equatorial area, Journal of the Meteorological Society of Japan. Ser. II 44 (1) (1966) 25–43.
  • [38] A. E. Gill, Some simple solutions for heat-induced tropical circulation, Quarterly Journal of the Royal Meteorological Society 106 (449) (1980) 447–462.
  • [39] G. K. Vallis, Geophysical fluid dynamics: whence, whither and why?, Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 472 (2192) (2016) 20160140.
  • [40] A. J. Majda, S. N. Stechmann, The skeleton of tropical intraseasonal oscillations, Proceedings of the National Academy of Sciences 106 (21) (2009) 8417–8422.
  • [41] M. A. Cane, E. Sarachik, D. Moore, E. Sarachik, The response of a linear baroclinic equatorial ocean to periodic forcing, Journal of Marine Research 39 (4) (1981) 651–693.
  • [42] A. Majda, Introduction to PDEs and Waves for the Atmosphere and Ocean, Vol. 9, American Mathematical Soc., 2003.
  • [43] D. Harrison, G. A. Vecchi, Westerly wind events in the tropical pacific, 1986–95, Journal of Climate 10 (12) (1997) 3131–3156.
  • [44] G. A. Vecchi, D. Harrison, Tropical Pacific sea surface temperature anomalies, El Niño, and equatorial westerly wind events, Journal of climate 13 (11) (2000) 1814–1830.
  • [45] E. Tziperman, L. Yu, Quantifying the dependence of westerly wind bursts on the large-scale tropical Pacific SST, Journal of Climate 20 (12) (2007) 2760–2768.
  • [46] S. Hu, A. V. Fedorov, Exceptionally strong easterly wind burst stalling El Niño of 2014, Proceedings of the National Academy of Sciences 113 (8) (2016) 2005–2010.
  • [47] A. F. Levine, M. J. McPhaden, How the July 2014 easterly wind burst gave the 2015–2016 El Niño a head start, Geophysical research letters 43 (12) (2016) 6503–6510.
  • [48] H. H. Hendon, M. C. Wheeler, C. Zhang, Seasonal dependence of the MJO–ENSO relationship, Journal of Climate 20 (3) (2007) 531–543.
  • [49] M. Puy, J. Vialard, M. Lengaigne, E. Guilyardi, Modulation of equatorial Pacific westerly/easterly wind events by the Madden–Julian oscillation and convectively-coupled Rossby waves, Climate Dynamics 46 (7-8) (2016) 2155–2178.
  • [50] C. W. Gardiner, et al., Handbook of stochastic methods, Vol. 3, springer Berlin, 1985.
  • [51] A. F. Levine, F.-F. Jin, Noise-induced instability in the ENSO recharge oscillator, Journal of the Atmospheric Sciences 67 (2) (2010) 529–542.
  • [52] N. Chen, F. Gilani, J. Harlim, A Bayesian machine learning algorithm for predicting ENSO using short observational time series, Geophysical Research Letters 48 (17) (2021) e2021GL093704.
  • [53] K. Ashok, S. K. Behera, S. A. Rao, H. Weng, T. Yamagata, El Niño Modoki and its possible teleconnection, Journal of Geophysical Research: Oceans 112 (C11).
  • [54] T. Lee, M. J. McPhaden, Increasing intensity of el Niño in the central-equatorial pacific, Geophysical Research Letters 37 (14).
  • [55] A. Capotondi, A. T. Wittenberg, M. Newman, E. Di Lorenzo, J.-Y. Yu, P. Braconnot, J. Cole, B. Dewitte, B. Giese, E. Guilyardi, et al., Understanding ENSO diversity, Bulletin of the American Meteorological Society 96 (6) (2015) 921–938.
  • [56] S. Thual, A. J. Majda, N. Chen, Statistical occurrence and mechanisms of the 2014–2016 delayed super El Niño captured by a simple dynamical model, Climate Dynamics 52 (3-4) (2019) 2351–2366.
  • [57] L. Chen, T. Li, B. Wang, L. Wang, Formation mechanism for 2015/16 super El Niño, Scientific reports 7 (1) (2017) 1–10.
  • [58] S.-I. An, S.-K. Kim, A. Timmermann, Fokker–Planck dynamics of the El Niño-Southern Oscillation, Scientific reports 10 (1) (2020) 1–11.
  • [59] S.-I. An, E. Tziperman, Y. M. Okumura, T. Li, ENSO irregularity and asymmetry, El Niño Southern Oscillation in a changing climate (2020) 153–172.
  • [60] A. F. Levine, F. F. Jin, A simple approach to quantifying the noise–enso interaction. part i: Deducing the state-dependency of the windstress forcing using monthly mean data, Climate Dynamics 48 (1-2) (2017) 1–18.
  • [61] L. T. Giorgini, W. Moon, N. Chen, J. Wettlaufer, Non-Gaussian stochastic dynamical model for the El Niño southern oscillation, Physical Review Research 4 (2) (2022) L022065.
  • [62] S. Zhao, F.-F. Jin, X. Long, M. A. Cane, On the breakdown of ENSO’s relationship with thermocline depth in the central-equatorial Pacific, Geophysical Research Letters 48 (9) (2021) e2020GL092335.