Articles | Volume 30, issue 3
https://doi.org/10.5194/npg-30-263-2023
https://doi.org/10.5194/npg-30-263-2023
Research article
 | 
06 Jul 2023
Research article |  | 06 Jul 2023

An adjoint-free algorithm for conditional nonlinear optimal perturbations (CNOPs) via sampling

Bin Shi and Guodong Sun
Abstract

In this paper, we propose a sampling algorithm based on state-of-the-art statistical machine learning techniques to obtain conditional nonlinear optimal perturbations (CNOPs), which is different from traditional (deterministic) optimization methods.1 Specifically, the traditional approach is unavailable in practice, which requires numerically computing the gradient (first-order information) such that the computation cost is expensive, since it needs a large number of times to run numerical models. However, the sampling approach directly reduces the gradient to the objective function value (zeroth-order information), which also avoids using the adjoint technique that is unusable for many atmosphere and ocean models and requires large amounts of storage. We show an intuitive analysis for the sampling algorithm from the law of large numbers and further present a Chernoff-type concentration inequality to rigorously characterize the degree to which the sample average probabilistically approximates the exact gradient. The experiments are implemented to obtain the CNOPs for two numerical models, the Burgers equation with small viscosity and the Lorenz-96 model. We demonstrate the CNOPs obtained with their spatial patterns, objective values, computation times, and nonlinear error growth. Compared with the performance of the three approaches, all the characters for quantifying the CNOPs are nearly consistent, while the computation time using the sampling approach with fewer samples is much shorter. In other words, the new sampling algorithm shortens the computation time to the utmost at the cost of losing little accuracy.

Dates
1 Introduction

One1 of the critical issues for weather and climate predictability is the short-term behavior of a predictive model with imperfect initial data. For assessing subsequent errors in forecasts, it is of vital importance to understand the model's sensitivity to errors in the initial data. Perhaps the simplest and most practical way is to estimate the likely uncertainty in the forecast by considering running it with initial data polluted by the most dangerous errors. The traditional approach is the normal mode method (Rayleigh1879; Lin1955), which is based on the linear stability analysis and has been used to understand and analyze the observed cyclonic waves and long waves of middle and high latitudes (Eady1949). However, atmospheric and oceanic models are often linearly unstable. Specifically, the transient growth of perturbations can still occur in the absence of growing normal modes (Farrell and Ioannou1996a, b). Therefore, the normal mode theory is generally unavailable for the prediction generated by atmospheric and oceanic flows. To achieve the goal of prediction on a low-order two-layer quasi-geostrophic model in a periodic channel, Lorenz (1965) first proposed a non-normal mode approach and simultaneously introduced some new concepts, i.e., the tangent linear model, adjoint model, singular values, and singular vectors. Then, Farrell (1982) used this linear approach to investigate the linear instability within finite time. During the last decade of the past century, such a linear approach had been widely used to identify the most dangerous perturbations of atmospheric and oceanic flows and also extended to explore error growth and predictability, such as patterns of the general atmospheric circulations (Buizza and Palmer1995) and the coupled ocean–atmosphere model of the El Niño–Southern Oscillation (ENSO) (Xue et al.1997a, b; Thompson1998; Samelson and Tziperman2001). Recently, the non-normal approach had been extended further to an oceanic study for investigating the predictability of the Atlantic meridional overturning circulation (Zanna et al.2011) and the Kuroshio path variations (Fujii et al.2008).

Both the approaches of normal and non-normal modes are based on the assumption of linearization, which means that the initial error must be so small that a tangent linear model can approximately quantify the error's growth. Besides, the complex nonlinear atmospheric and oceanic processes have not yet been well considered in the literature. To overcome this limitation, Mu (2000) proposed a nonlinear non-normal mode approach with the introduction of two new concepts, nonlinear singular values and nonlinear singular vectors. Mu and Wang (2001) then used it to successfully capture the local fastest-growing perturbations for a 2D quasi-geostrophic model. However, several disadvantages still exist, such as unavailability in practice and unreasonable physics of the large norm for local fastest-growing perturbations. Starting from the perspective of nonlinear programming, Mu et al. (2003) proposed an innovative approach, named conditional nonlinear optimal perturbation (CNOP), to explore the optimal perturbation that can fully consider the nonlinear effect without any assumption of linear approximation. Generally, the CNOP approach captures initial perturbations with maximal nonlinear evolution given by a reasonable constraint in physics. Currently, the CNOP approach as a powerful tool has been widely used to investigate the fastest-growing initial error in the prediction of an atmospheric and oceanic event and to reveal some related mechanisms, such as the stability of the thermohaline circulation (Mu et al.2004; Zu et al.2016), the predictability of ENSO (Duan et al.2009; Duan and Hu2016) and the Kuroshio path variations (Wang and Mu2015), the parameter sensitivity of the land surface processes (Sun and Mu2017), and typhoon-targeted observations (Mu et al.2009; Qin and Mu2012). Some more details are shown in Wang et al. (2020) and more perspectives on general fluid dynamics in Kerswell (2018).

The primary goal of obtaining the CNOPs is to efficiently and effectively implement nonlinear programming, mainly including the spectral projected gradient (SPG) method (Birgin et al.2000), sequential quadratic programming (SQP) (Barclay et al.1998), and the limited memory Broyden–Fletcher–Goldfarb–Shanno (BFGS) algorithm (Liu and Nocedal1989) in practice. The gradient-based optimization algorithms are also adopted in the field of fluid dynamics to capture the minimal finite amplitude disturbance that triggers the transition to turbulence in shear flows (Pringle and Kerswell2010) and the maximal perturbations of the disturbance energy gain in a 2D isolated counter-rotating vortex pair (Navrose et al.2018). For the CNOP, the objective function of the initial perturbations in a black-box model is obtained by the error growth of a nonlinear differential equation. Therefore, the essential difficulty here is how to efficiently compute the gradient (first-order information). Generally for an earth system model or an atmosphere–ocean general circulation model, computing the gradient directly from the definition of numerical methods is unavailable, since it requires plenty of runs of the nonlinear model. Perhaps the most popular and practical way to numerically approximate the gradient is the adjoint technique, which is based on deriving the tangent linear model and the adjoint model (Kalnay2003). Once we can distill out the adjoint model, computing the gradient at the cost of massive storage space to save the basic state becomes available. In other words, the adjoint-based method takes a large amount of storage space to reduce computation time significantly. However, the adjoint model is unusable for many atmospheric and oceanic models, since it is hard to develop, especially for the coupled ocean–atmosphere models. Still, the adjoint-based method can only deal with a smooth case. Wang and Tan (2010) proposed the ensemble-based methods, which introduces the classical techniques of empirical orthogonal function (EOF) decomposition widely used in atmospheric science and oceanography. Specifically, it takes some principal modes of the EOF decomposition to approximate the tangent linear matrix. However, the colossal memory and repeated calculations occurring in the adjoint-based method still exist (Wang and Tan2010; Chen et al.2015). In addition, the intelligent optimization methods (Zheng et al.2017; Yuan et al.2015) are unavailable on the high-dimension problem (Wang et al.2020). All of the traditional (deterministic) optimization methods above cannot guarantee finding an optimal solution.

To overcome the limitations of the adjoint-based method described above, we start to take consideration from the perspective of stochastic optimization methods, which as the workhorse have powered recent developments in statistical machine learning (Bottou et al.2018). In this paper, we use the stochastic derivative-free method proposed by Nemirovski and Yudin (1983, Sect. 9.3.2) that takes a basis on the simple high-dimensional divergence theorem (i.e., Stokes' theorem). With this stochastic derivative-free method, the gradient can be reduced to the objective function value in terms of expectation. For the numerical implementation as an algorithm, we take the sample average to approximate the expectation. Based on the law of large numbers, we provide a concentration estimate for the gradient by the general Chernoff-type inequality. This paper is organized as follows. The basic description of the CNOP settings and the proposed sampling algorithm are given in Sects. 2 and 3, respectively. We then perform the preliminary numerical test for two numerical models, the simple Burgers equation with small viscosity and the Lorenz-96 model in Sect. 4. A summary and discussion are included in Sect. 5.

2 The Basic CNOP settings

In this section, we provide a brief description of the CNOP approach. It should be noted that the CNOP approach has been extended to investigate the influences of other errors, i.e., parameter errors and boundary condition errors, on atmospheric and oceanic models (Mu and Wang2017) beyond the original intention of CNOPs exploring the impact of initial errors. We only focus on the initial perturbations in this study.

Let Ω be a region in d with ∂Ω as its boundary. An atmospheric or oceanic model is governed by the following differential equation as

(1) U t = F ( U , P ) U | t = 0 = U 0 U | Ω = G ,

where U is the reference state in the configuration space; P is the set of model parameters; F is a nonlinear operator; and U0 and G are the initial reference state and the boundary condition, respectively. Without loss of generality, we note gt(⋅) to be the reference state evolving with time U(t;) in the configuration space. Thus, given any initial state U0, we can obtain that the reference state at time T is gT(U0)=U(T;U0). If we consider the initial state U0+u0 as the perturbation of U0, then the reference state at time T is given by gT(U0+u0)=U(T;U0+u0) .

With both the reference states at time T, gT(U0) and gT(U0+u0), the objective function of the initial perturbation u0 based on the initial condition U0 is

(2) J ( u 0 ; U 0 ) = g T ( U 0 + u 0 ) - g T ( U 0 ) 2 , 2

and2 then the CNOP formulated as the constrained optimization problem is

(3) max u 0 δ J ( u 0 ; U 0 ) .

Both the objective function (Eq. 2) and the optimization problem (Eq. 3) come directly from the theoretical model (Eq. 1). When we take the numerical computation, the properties of the two objects above, Eqs. (2) and (3), probably become different. Here, it is necessary to mention some similarities and dissimilarities between the theoretical model (Eq. 1) and its numerical implementation. If the model (Eq. 1) is a system of ordinary differential equations, then it is finite dimensional, and so there are no other differences between the theoretical model (Eq. 1) and its numerical implementation except some numerical errors. However, if the model (Eq. 1) is a partial differential equation, then it is infinite dimensional. When we implement it numerically, the dimension is reduced to be finite for both the objective function (Eq. 2) and the optimization problem (Eq. 3). At last, we conclude this section with the notation J(u0) shortened J(u0;U0) afterward for convenience.

3 Sample-based algorithm

In this section, we first describe the basic idea of the sampling algorithm. Then, it is shown for comparison with the baseline algorithms in numerical implementation. Finally, we conclude this section with a rigorous Chernoff-type concentration inequality to characterize the degree to which the sample average probabilistically approximates the exact gradient. The detailed proof is postponed to Appendix A.

The key idea for us to consider the sampling algorithm is based on the high-dimensional Stokes' theorem, which reduces the gradient in the unit ball to the objective value on the unit sphere in terms of the expectation. Let 𝔹d be the unit ball in d and v0∼Unif(𝔹d) a random variable v0 following the uniform distribution in the unit ball 𝔹d. Given a small real ϵ>0, we can define the expectation of the objective function J in the ball with the center u0 and the radius ϵ as

(4) J ^ ( u 0 ) = E v 0 B d [ J ( u 0 + ϵ v 0 ) ] .

In other words, the objective function J is required to define in the ball B(0;δ+ϵ)={u0Rd:u0ϵ+δ}. Also, we find that J^(u0) is approximate to J(u0); that is, J^(u0)J(u0). If the gradient J exists in the ball B(0;δ+ϵ), the fact that the expectation of v0 is zero tells us that the error of the objective value is estimated as

J(u0)-J^(u0)=O(ϵ2).

Before proceeding to the next, we note the unit sphere as Sd-1=Bd. With the representation of J^(u0) in Eq. (4), we can obtain the gradient J^(u0) directly from the function value J by the high-dimensional Stokes' theorem as

(5) J ^ ( u 0 ) = E v 0 B d [ J ( u 0 + ϵ v 0 ) ] = d ϵ E v 0 S d - 1 [ J ( u 0 + ϵ v 0 ) v 0 ,

where v0Unif(Sd-1) in the last equality. Similarly, J^(u0) is approximate to J(u0); that is, J^(u0)J(u0). If the gradient J exists in the ball B(0;δ+ϵ), we can show that the error of the gradient is estimated as

(6) J ^ ( u 0 ) - J ( u 0 ) = O ( d ϵ ) .

The rigorous description and proof are shown in Appendix A (Lemma A1 with its proof).

In the numerical computation, we obtain the approximate gradient, J^(u0), via the sampling as

(7) d ϵ E v 0 S d - 1 [ J ( u 0 + ϵ v 0 ) v 0 ] d n ϵ i = 1 n J ( u 0 + ϵ v 0 , i ) v 0 , i ,

where v0,iUnif(Sd-1) and (i=1,,n) are the independent random variables following the identical uniform distribution on 𝕊d−1. Since the expectation of the random variable v0 on the unit sphere 𝕊d−1, we generally take the following way with better performance in practice as

(8) d ϵ E v 0 S d - 1 [ J ( u 0 + ϵ v 0 ) v 0 ] = d ϵ E v 0 S d - 1 [ ( J ( u 0 + ϵ v 0 ) - J ( u 0 ) ) v 0 ] d n ϵ i = 1 n ( J ( u 0 + ϵ v 0 , i ) - J ( u 0 ) ) v 0 , i ,

where v0,iUnif(Sd-1) and (i=1,,n) are independent. From Eq. (8), n is the number of samples and d is the dimension. Generally in practice, the number of samples is far less than the dimension, nd. Hence, the times to run the numerical model is n+1d+1, which is the times to run the numerical model via the definition of the numerical method as

J(u0)u0,iJ(u0+ϵei)-J(u0)ϵ,

where i=1,,d. For the adjoint method, the gradient is numerically computed as

J(u0)MMu0MgT(U0+u0),

where M is a product of some tangent linear models. Practically, the adjoint model, MT, is hard to develop. In addition. we cannot obtain the tangent linear model for the coupled ocean–atmosphere models.

Next, we provide a simple but intuitive analysis of the convergence in probability for the samples in practice. With the representation of J^(u0) in Eq. (5), the weak law of large numbers states that the sample average (Eq. 7) converges in probability toward the expected value; that is, for any t>0, we have

Prdnϵi=1nJ(u0+ϵv0,i)v0,i-J^(u0)t0,withn.

Combined with the error estimate of gradient (Eq. 6), if t is assumed to be larger than Ω(dϵ) (i.e., there exists a constant τ>0 such that t>τdϵ), then the probability that the sample average approximates to J(u0) satisfies

Prdnϵi=1nJ(u0+ϵv0,i)v0,i-J(u0)t-Ω(dϵ)0,withn.

Finally, we conclude the section with the rigorous Chernoff-type bound in probability for the simple but intuitive analysis above with the following theorem. The rigorous proof is shown in Appendix A with Lemma A2 and Lemma A3 proposed.

Theorem 1.

If J is continuously differentiable and satisfies the gradient Lipschitz condition, i.e., for any u0,1,u0,2B(0,δ), there exists a constant L>0 such that the following inequality holds as

J(u0,1)-J(u0,2)Lu0,1-u0,2.

For any t>Ldϵ/2, there exists a constant C>0 such that the samples satisfy the concentration inequality as

Pr(dnϵi=1nJ(u0+ϵv0,i)v0,i-J(u0)t-Ldϵ2)2exp(-Cnt2).
4 Numerical model and experiments

In this section, we perform several experiments to compare the proposed sampling algorithm with the baseline algorithms for two numerical models, the Burgers equation with small viscosity and the Lorenz-96 model. After the CNOP was first proposed in Mu et al. (2003), plenty of methods, adjoint based or adjoint free, have been introduced to compute the CNOPs (Wang and Tan2010; Chen et al.2015; Zheng et al.2017; Yuan et al.2015). However, some essential difficulties have still not been overcome. Taking the classical adjoint technique for example, the massive storage space and unusability in many atmospheric and oceanic modes are the two insurmountable points. In this study, different from traditional (deterministic) optimization methods above, we obtain the approximate gradient by sampling the objective values introduced in Sect. 3. Then, we use the second spectral projected gradient method (SPG2) proposed in Birgin et al. (2000) to compute the CNOPs.

Table 1The objective values of CNOPs and the percentage over that computed by the definition method. Bold emphasizes the high efficiency of the sampling method.

Download Print Version | Download XLSX

https://npg.copernicus.org/articles/30/263/2023/npg-30-263-2023-f01

Figure 1Spatial distributions of CNOPs (unit: m s−1). Prediction time: on the top is T= 30 s, and on the bottom is T= 60 s.

Download

4.1 The Burgers equation with small viscosity

We first consider a simple theoretical model, the Burgers equation with small viscosity under the Dirichlet condition. It should be noted here that we adopt the internal units meters and seconds. The reference state U evolves nonlinearly with time as

(9) U t + U U x = γ 2 U x 2 , ( x , t ) [ 0 , L ] × [ 0 , T ] U ( 0 , t ) = U ( L , t ) = 0 , t [ 0 , T ] U ( x , 0 ) = sin 2 π x L , x [ 0 , T ] ,

where γ= 0.005 m2 s−1 and L= 100 m. We use the DuFort–Frankel scheme or the leapfrog scheme (i.e., the central finite difference scheme in both the temporal and spatial directions) to numerically solve the viscous Burgers equation above (Eq. 9), with Δx= 1 m as the spatial grid size (d=101) and Δt= 1 s as the time step. The objective function J(u0) used for optimization (Eq. 2) can be rewritten in the form of the perturbation's norm square as

J(u0)=u(T)2=i=1dui(T)2.

The constraint parameter is set to be δ= 8 × 10−4m s−1 such that the initial perturbation satisfies u0δ= 8 × 10−4m s−1. We note the numerical gradient computed directly as the definition method, where the step size for the difference is set to be h=10-8. Together with the adjoint methods, we set them as the baseline algorithms. For the sampling algorithm, we set the parameter ϵ= 10−8 in Eq. (4), the expectation of the objective function. The time T in the objective function (Eq. 2) is named as the prediction time. We take the two groups of numerical experiments according to the prediction time, T= 30 s and T= 60 s, to calculate the CNOPs by the baseline algorithms and the sampling method. And then, we compare their performances to show the superiority of the sampling method.

The spatial distributions of the CNOPs computed by the baseline algorithms and the sampling method are shown in Fig. 1, where we can find the change of the CNOPs' spatial pattern for the Burgers equation with small viscosity as follows:

  1. The spatial pattern of the CNOPs computed by two baseline algorithms, the definition method and the adjoint method, are nearly identical.

  2. Based on the spatial pattern of the CNOPs computed by two baseline algorithms, there are some fluctuating errors for the sampling method.

  3. When the number of samples increase from n=5 to n=15, the fluctuating errors in the spatial patter are reduced.

Table 2Comparison of computation times (unit: s). Run on Matlab2022a with Intel® Core™ i9-10900 CPU of 2.80 GHz. Bold emphasizes the high efficiency of the sampling method.

Download Print Version | Download XLSX

https://npg.copernicus.org/articles/30/263/2023/npg-30-263-2023-f02

Figure 2Nonlinear evolution behavior of the CNOPs in terms of the norm square.

Download

We have figured out the spatial distributions of the CNOPs in Fig. 1. In other words, we have qualitatively characterized the CNOPs. However, we still need some quantities to measure the CNOPs' performance. Here, we use the objective value of the CNOP as the quantity. The objective values corresponding to the spatial patterns in Fig. 1 are shown in Table 1. To show them clearly, we make them over that computed by the definition method with the percentage representation, which is also shown in Table 1. For the Burgers equation with small viscosity, we can also find that the objective values by the adjoint method over that by the definition method is 100 % for both the cases, T= 30 s and T= 60 s, respectively; the objective values by the sampling method are all more than 90 %; and when we increase the number of samples from n=5 to n=15, the percentage increases from 94.95 % to 96.75 % for the case T= 30 s and from 95.46 % to 97.57 % for the case T= 60 s. The objective values of the CNOPs in Table 1 quantitatively echo the performances of spatial patterns in Fig. 1.

Next, we show the computation times to obtain the CNOPs by the baseline algorithms and the sampling method in Table 2. For the Burgers equation with small viscosity, the computation time taken by the adjoint method is far less than that directly by the definition method for the two cases, T= 30 s and T= 60 s. When we implement the sampling method, the computation time using n=15 samples is almost the same as that taken directly by the definition method. If we reduce the number of samples from n=15 to n=5, the computation time is shortened by more than half.

Finally, we describe the nonlinear evolution behavior of the CNOPs in terms of norm squares u(t)2 computed by the baseline algorithms and the sampling method in Fig. 2. We can find that there exists a fixed turning-time point, approximately t=20 s for the prediction time T= 30 s and approximately t= 50 s for T= 60 s. In the beginning, the nonlinear growth of the CNOPs is very slow. When the evolving time comes across the fixed turning-time point, the perturbations start to proliferate. Figure 2 shows the nonlinear evolution behaviors of the CNOPs computed by all the algorithms above are almost consistent but do not provide any tiny difference between the baseline algorithms and the sampling method. So we further show the nonlinear evolution behavior of the CNOPs in terms of the difference Δu(t)2 and relative difference Δu(t)2/u(t)2 based on the definition method in Fig. 3. There is no difference or relative difference in the nonlinear error growth between the two baseline algorithms. The top two graphs in Fig. 3 show that the differences do not grow fast until the time comes across the turning-time point. When we reduce the number of samples, the difference enlarges gradually, with the maximum around 6 × 10−7m2 s−2 for T= 30 s and 0.12 m2 s−2 for T= 60 s. However, the differences are very small compared with the nonlinear growth of the CNOPs themselves, which is shown by the relative difference in the bottom two graphs of Fig. 3. In addition, some numerical errors exist around t= 11 s for the relative difference and decrease with increasing the number of samples.

https://npg.copernicus.org/articles/30/263/2023/npg-30-263-2023-f03

Figure 3Nonlinear evolution behavior of the CNOPs in terms of the difference and relative difference of the norm square.

Download

The Burgers equation with small viscosity is a partial differential equation, which is an infinite-dimensional dynamical system. In the numerical implementation, it corresponds to the high-dimensional case. Taking all the performances with different test quantities into account, i.e., spatial structures, objective values, computation times, and nonlinear error growth, we conclude that the adjoint method obtains almost the total information and saves much computation time simultaneously; the sampling method with n=15 samples drops a few accuracies and loses little information but shares nearly the same computation time; and when we reduce the number of samples from n=15 to n=5, we can obtain about 95 % of information as the baseline algorithms, but the computation time is reduced by more than half. The cause for the phenomenon described above is perhaps due to the high-dimensional property.

4.2 The Lorenz-96 model

Next, we consider the Lorenz-96 model, one of the most classical and idealized models, which is designed to study fundamental issues regarding the predictability of the atmosphere and weather forecasting (Lorenz2006; Lorenz and Emanuel1998). In recent 2 decades, the Lorenz-96 model has been widely applied in data assimilation and predictability (Ott et al.2004; Trevisan and Palatella2011; De Leeuw et al.2018) to studies in spatiotemporal chaos (Pazó et al.2008). The Lorenz-96 model is also used to investigate the predictability of extreme amplitudes of traveling waves (Sterk and van Kekem2017), which points out that it depends on the dynamical regime of the model.

With a cyclic permutation of the variables as {xi}i=1N satisfying x-1=xN-1, and x0=xN, x1=xN+1, the governing system of equations for the Lorenz-96 model is described as

(10) d x i d t = ( x i + 1 - x i - 2 ) x i - 1 advection - x i damping + F external forcing ,

where the system is nondimensional. The physical mechanism considers that the total energy is conserved by the advection, decreased by the damping and kept away from zero by the external forcing. The variables xi (i=1,,N) can be interpreted as values of some atmospheric quantity (e.g., temperature, pressure, or vorticity) measured along a circle of constant latitude of the earth (Lorenz2006). The Lorenz-96 model (Eq. 10) can also describe waves in the atmosphere. Specifically, Lorenz (2006) observed that the waves slowly propagate westward toward decreasing i for F>0 that are sufficiently large.

In this study, we use the classical fourth-order Runge–Kutta method to numerically solve the Lorenz-96 model (Eq. 10). The two parameters are set as N=40 and F=8, respectively. The spatial distributions of the CNOPs computed by the baseline algorithms and the sampling method are shown in Fig. 4. Unlike the three dominant characters described above for the Burgers equation with small viscosity, the spatial distributions of the CNOPs computed by all four algorithms are almost consistent except for some little fluctuations for the Lorenz-96 model. Similarly, we provide a display for the objective values of the CNOPs computed by the baseline algorithms and the sampling method and the percentage over that computed by the definition method in Table 3. We find that the percentage of the objective value computed by the adjoint method is only 92.35 %, less than that by the sampling method for the Lorenz-96 model. In addition, the difference between the number of samples n=5 and n=15 is only 0.57 % in the percentage of the objective value. In other words, we can obtain about 95 % of the total information by taking the sampling algorithm only using n=5 samples.

https://npg.copernicus.org/articles/30/263/2023/npg-30-263-2023-f04

Figure 4Spatial distributions of CNOPs.

Download

Table 3The objective values of CNOPs and the percentage over that computed by the definition method. Bold emphasizes the high efficiency of the sampling method.

Download Print Version | Download XLSX

Similarly, we show the computation times to obtain the CNOPs by the baseline algorithms and the sampling method in Table 4. For the adjoint method, different from the Burgers equation with small viscosity, the time to compute the CNOP by the adjoint method is almost twice that used by the definition method for the Lorenz-96 model. However, the computation time that the sampling method uses is less than one-third of what the definition method uses. When we reduce the number of samples from n=15 to n=5, the computation time decreases by more than 0.1 s. As a result, the sampling method only using n=5 samples saves much computation time to obtain the CNOP for the Lorenz-96 model.

Table 4Comparison of computation times. Run on Matlab2022a with Intel® Core™ i9-10900 CPU of 2.80 GHz. Bold emphasizes the high efficiency of the sampling method.

Download Print Version | Download XLSX

https://npg.copernicus.org/articles/30/263/2023/npg-30-263-2023-f05

Figure 5Nonlinear evolution behavior of the CNOPs in terms of the norm square.

Download

https://npg.copernicus.org/articles/30/263/2023/npg-30-263-2023-f06

Figure 6Nonlinear evolution behavior of the CNOPs in terms of the difference and relative difference of the norm square.

Download

Finally, we demonstrate the nonlinear evolution behavior of the CNOPs in terms of norm squares x(t)2 computed by the baseline algorithms and the sampling method in Fig. 5. Recall Fig. 2 for the Burgers equation with small viscosity; the norm squares of the CNOPs have almost no growth until the turning-time point and then proliferate. Unlikely, the norm squares of the CNOPs almost grow linearly for the Lorenz-96 model without any turning-time point. Similarly, since the four nonlinear evolution curves of norm squares almost coincide in Fig. 5, we cannot find any tiny difference in the nonlinear growth of the CNOPs between the baseline algorithms and the sampling method. So we still need to observe the nonlinear evolution behavior of the CNOPs in terms of the difference Δx(t)2, which is shown in the left panel of Fig. 6. In the initial stage, the three nonlinear evolution curves share the same growth behavior, with the maximum amplitude being the one by implementing the sampling method with n=5 samples. Afterward, the error's amplitude decreases for the sampling method with n=5 samples, and the two curves by the adjoint method and the sampling method with n=5 samples are similar, with the larger amplitude being the one by the adjoint method, which achieves the maximum around 0.3. Indeed, the differences are very small compared with the nonlinear growth of the CNOPs themselves, which is shown by the relative difference Δx(t)2/x(t)2 in the right panel of Fig. 6. We can find that the three curves of the relative difference share the same nonlinear evolution behavior, with the sampling method of different numbers of samples on both sides and the adjoint method in between. When the number of samples is reduced, the amplitude of the relative difference decreases. In addition, the order of the relative difference's magnitude is 10−3, which is so tiny that there are no essential differences.

Although the dimension of the Lorenz-96 model is not very large due to being composed of a finite number of ordinary differential equations, it possesses strongly nonlinear characters. Unlike the Burgers equation with small viscosity, the adjoint method does not work well for the Lorenz-96 model, which spends more computation time and obtains less percentage of the total information. The sampling method performs more advantages in the computation, saving far more computation time and obtaining more information. However, the performance in reducing the number of samples from n=15 to n=5 is not obvious. Perhaps this is due to the characters of the Lorenz-96 model of strong nonlinearity and low dimension.

5 Summary and discussion

In this paper, we introduce a sampling algorithm to compute the CNOPs based on the state-of-the-art statistical machine learning techniques. The theoretical guidance comes from the high-dimensional Stokes' theorem and the law of large numbers. We derive a Chernoff-type concentration inequality to rigorously characterize the degree to which the sample average probabilistically approximates the exact gradient. We show the advantages of the sampling method by comparison with the performance of the baseline algorithms, e.g., the definition method and the adjoint method. If there exists the adjoint model, the computation time is reduced significantly with the exchange of much storage space. However, the adjoint model is unusable for the complex atmospheric and oceanic model in practice.

For the numerical tests, we choose two simple but representative models, the Burgers equation with small viscosity and the Lorenz-96 model. The Burgers equation with small viscosity is one of the simplest nonlinear partial differential equations simplified from the Navier–Stokes equation, which holds a high-dimensional property. The Lorenz-96 model is a low-dimensional dynamical system with strong nonlinearity. For the numerical performance of a partial differential equation, the Burgers equation with small viscosity, we find that the adjoint method performs very well and saves much computation time; the sampling method can share nearly the same computation time with the adjoint method with dropping a few accuracies by adjusting the number of samples; and the computation time can be shortened more by reducing the number of samples further with the nearly consistent performance. For the numerical performance of a low-dimensional and strong nonlinear dynamical system, the Lorenz-96 model, we find that the adjoint method takes underperformance, but the sampling method fully occupies the dominant position, regardless of saving the computation time and performing the CNOPs in terms of the spatial pattern, the objective value, and the nonlinear growth. Still, unlike the Burgers equation with small viscosity, the performance is not obvious for reducing the number of samples for the Lorenz-96 model. Based on the comparison above, we propose a possible conclusion that the sampling method probably works very well for an atmospheric or oceanic model in practice, which is a partial differential equation with strong nonlinearity. Perhaps the high efficiency of the sampling method performs more dominantly, and the computation time is shortened obviously by reducing the number of samples.

Currently, the CNOP method has been widely applied to predictability in meteorology and oceanography. For the nonlinear multiscale interaction (NMI) model (Luo et al.2014, 2019), an atmospheric blocking model which successfully characterizes the life cycle of the dynamic blocking from onset to decay, the CNOP method has been used to investigate the sensitivity on initial perturbations and the impact of the westerly background wind (Shi et al.2022). However, it is still very challenging to compute the CNOP for a realistic earth system model, such as the Community Earth System Model (CESM) (Wang et al.2020). Many difficulties still exist, even for a high-regional resolution model, such as the Weather Research and Forecasting (WRF) model, which is used widely in operational forecasting (Yu et al.2017). Based on increasingly reliable models, we now comment on some extensions for further research to compute and investigate the CNOPs on the more complex models by the sampling method, regardless of theoretical or practical. An idealized ocean–atmosphere coupling model, the Zebiak–Cane (ZC) model (Zebiak and Cane1987), might characterize the oscillatory behavior of ENSO in amplitude and period based on oceanic wave dynamics. Mu et al. (2007) computed the CNOP of the ZC model by use of its adjoint model to study the spring predictability barrier for El Niño events. In addition, Mu et al. (2009) also computed the CNOP of the PSU/NCAR (Pennsylvania State University/National Center for Atmospheric Research) mesoscale model (i.e., the MM5 model) using its adjoint model to explore the predictability of tropical cyclones. It looks very interesting and practical to test the validity of the sampling algorithm to calculate the CNOPs on the two more realistic numerical models, the ZC model and the MM5 model, as well as the idealized NMI model. For an earth system model or atmosphere–ocean general circulation models (AOGCMs), it is often impractical to obtain the adjoint model, so the sampling method provides a probable way of computing the CNOPs to investigate its predictability. In addition, it becomes possible for us to use 4D-Var data assimilation on a coupled climate system model when the sampling method is introduced. Also, it is thrilling to implement the sampling method in the Flexible Global Ocean-Atmosphere-Land System (FGOALS)-s2 (Wu et al.2018) for decadal climate prediction.

Appendix A: Proof of Theorem 1

Lemma A1. If the expectation J^(u0) is given by Eq. (4), then the expression (Eq. 5) is satisfied. Also, under the same assumption of Theorem 1, the difference between the expectation of objective value and itself can be estimated as

(A1) J ^ ( u 0 ) - J ( u 0 ) L ϵ 2 2 ,

and the difference between the expectation of gradient and itself can be estimated as

(A2) J ^ ( u 0 ) - J ( u 0 ) L d ϵ 2 .

Proof of Lemma A1. First, with the definition of J^(u0), we show the proof of Eq. (5), the equivalent representation of the gradient J^(u0).

  • For d=1, the gradient of the expectation J^ about u0 can be computed as

    dJ^(u0)du0=ddu012-11J(u0+ϵv0)dv0=12-11dJ(u0+ϵv0)ϵdv0dv0=J(u0+ϵ)-J(u0-ϵ)2ϵ.
  • For the case of d≥2, we assume that a∈ℝd is an arbitrary vector. Then, the gradient J^(u0) satisfies the following equality as

    aJ^(u0)=v0Bdau0J(u0+ϵv0)dV=1ϵv0Bdv0(J(u0+ϵv0)a)dV=1ϵv0Sd-1J(u0,k+ϵv0)av0dS=a1ϵv0Sd-1J(u0+ϵv0)v0dS.

    Because the vector a is arbitrary, we can obtain the following equality:

    v0BdJ(u0+ϵv0)dV=1ϵv0Sd-1J(u0+ϵv0)v0dS.

    Since the ratio of the surface area and the volume of the unit ball 𝔹d is d, the equivalent representation of the gradient (Eq. 5) is satisfied.

If the objective function J is continuously differentiable and satisfies the gradient Lipschitz condition, we can obtain the following inequality as

|J(u0+ϵv0)-J(u0)-ϵJ(u0),v0|Lϵ22v02.

Because v0BdJ(u0),v0dV=0, the estimate (Eq. A1) is obtained directly.

  • For any ij{1,,d}, since v0,i and v0,j are uncorrelated, we have

    v0Sd-1v0,iv0,jdS=0.
  • For any i=j{1,,d}, we have

    v0Sd-1v0,i2dS=1dv0Sd-1i=1dv0,i2dS=1dv0Sd-1dS.

Since v0 is a row vector, we derive the following equality as

Ev0Sd-1[v0Tv0]=1dI.

Hence, we obtain the equivalent representation of the gradient J(u0) as

J(u0)=dϵEv0Sd-1ϵJ(u0),v0v0.

Finally, since v0Unif(Sd-1), then Ev0Sd-1[v0]=0. Hence, we estimate the difference between the expectation of gradient and itself as

J^(u0)-J(u0)dϵEv0Sd-1J(u0+ϵv0)-J(u0)v0-dϵEv0Sd-1ϵJ(u0),v0v0dϵEv0Sd-1[J(u0+ϵv0)-J(u0)-ϵJ(u0),v0v0]Ldϵ2,

where the last inequality follows the gradient Lipschitz condition.

Considering any ϵ>0, to proceed with the concentration inequality, we still need to know that the random variable J(u0+ϵv0) for v0Unif(Sd-1) is sub-Gaussian. Thus, we first introduce the following lemma. Lemma A2 (Proposition 2.5.2 in Vershynin2018). Let X be a random variable. If there exist two constants K1,K2>0 such that the moment generating function of X2 is bounded:

EexpX2K12K2,

then the random variable X is sub-Gaussian.

Because J(u0+ϵv0) is bounded on 𝕊d−1, exp(J(u0+ϵv0)2/K12) is integrable on 𝕊d−1 for any K1>0; i.e., there exists a constant K2>0 such that

Ev0Sd-1expJ(u0+ϵv0)2K12K2.

With Lemma A2, the random variable J(u0+ϵv0) is sub-Gaussian. Therefore, for any fixed vector v0Sd-1, we know the random variable J(u0+ϵv0)v0,v0 is sub-Gaussian. We now introduce the following lemma to proceed with the concentration inequality.

Lemma A3 (Theorem 2.6.3 in Vershynin2018). Let X1,,Xn be independent, mean zero, sub-Gaussian random variables and a=(a1,,an)Rn. Then, for every t≥0, we have

Pri=1naiXit2exp-ct2K2a2,

where K=max1inXiψ2.3

Combined with Lemma A1 and Lemma A3, we can obtain the concentration inequality for the samples as

Pr|dnϵi=1nJ(u0+ϵv0,i)v0,iv0-J^(u0),v0|t2exp-cnt2K2,

where v0 is any unit vector on 𝕊d−1. Thus we can proceed with the concentration estimate by the Cauchy–Schwarz inequality as

Prdnϵi=1nJ(u0+ϵv0,i)v0,i-J^(u0)t2exp-cnt2K2.

Based on the triangle inequality, we can proceed with the concentration inequality with the estimate of the difference between the expectation of objective value and itself (Eq. A2) as

Prdnϵi=1nJ(u0+ϵv0,i)v0,i-J(u0)t-Ldϵ22exp-cnt2K2,

for any t>Ldϵ/2. Taking C=c/K2, we complete the proof of Theorem 1.

Code availability

The codes that support the findings of this study are available from the corresponding author, Bin Shi, upon reasonable request.

Data availability

No data sets were used in this article.

Author contributions

BS constructed the basic idea of this paper, derived all formulas and the proofs, coded the sampling method in MATLAB to show all the figures, and wrote the paper. GS joined the discussions of this paper and provided some suggestions. All the authors contributed to the writing and reviewing of the paper.

Competing interests

The contact author has declared that none of the authors has any competing interests.

Disclaimer

Publisher's note: Copernicus Publications remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Acknowledgements

We are indebted to Mu Mu for seriously reading an earlier version of this paper and providing his suggestions about this theoretical study. Bin Shi would also like to thank Ya-xiang Yuan, Ping Zhang, and Yu-hong Dai for their encouragement to understand and analyze the nonlinear phenomena in nature from the perspective of optimization in the early stages of this project. This work was supported by grant no. 12241105 of NSFC and grant no. YSBR-034 of CAS.

Financial support

This research has been supported by the National Natural Science Foundation of China (grant no. 12241105) and the Chinese Academy of Sciences (grant no. YSBR-034).

Review statement

This paper was edited by Stefano Pierini and reviewed by Stéphane Vannitsem and two anonymous referees.

References

Barclay, A., Gill, P. E., and Ben Rosen, J.: SQP methods and their application to numerical optimal control, in: Variational calculus, optimal control and applications, edited by: Schmidt, W. H., Heier, K., Bittner, L., and Bulirsch, R., Birkhäuser, Basel, 207–222, Print ISBN 978-3-0348-9780-8, Online ISBN 978-3-0348-8802-8, https://doi.org/10.1007/978-3-0348-8802-8_21, 1998. a

Birgin, E. G., Martínez, J. M., and Raydan, M.: Nonmonotone spectral projected gradient methods on convex sets, SIAM J. Optimiz., 10, 1196–1211, 2000. a, b

Bottou, L., Curtis, F. E., and Nocedal, J.: Optimization methods for large-scale machine learning, SIAM Rev., 60, 223–311, 2018. a

Buizza, R. and Palmer, T. N.: The singular-vector structure of the atmospheric global circulation, J. Atmos. Sci., 52, 1434–1456, 1995. a

Chen, L., Duan, W., and Xu, H.: A SVD-based ensemble projection algorithm for calculating the conditional nonlinear optimal perturbation, Sci. China Earth Sci., 58, 385–394, 2015. a, b

De Leeuw, B., Dubinkina, S., Frank, J., Steyer, A., Tu, X., and Vleck, E. V.: Projected shadowing-based data assimilation, SIAM J. Appl. Dyn. Syst., 17, 2446–2477, 2018. a

Duan, W. and Hu, J.: The initial errors that induce a significant “spring predictability barrier” for El Niño events and their implications for target observation: Results from an earth system model, Clim. Dynam., 46, 3599–3615, 2016. a

Duan, W., Liu, X., Zhu, K., and Mu, M.: Exploring the initial errors that cause a significant “spring predictability barrier” for El Niño events, J. Geophys. Res.-Oceans, 114, C04022, https://doi.org/10.1029/2008JC004925, 2009. a

Eady, E. T.: Long waves and cyclone waves, Tellus, 1, 33–52, 1949. a

Farrell, B. F.: The initial growth of disturbances in a baroclinic flow, J. Atmos. Sci., 39, 1663–1686, 1982. a

Farrell, B. F. and Ioannou, P. J.: Generalized stability theory. Part I: Autonomous operators, J. Atmos. Sci., 53, 2025–2040, 1996a. a

Farrell, B. F. and Ioannou, P. J.: Generalized stability theory. Part II: Nonautonomous operators, J. Atmos. Sci., 53, 2041–2053, 1996b. a

Fujii, Y., Tsujino, H., Usui, N., Nakano, H., and Kamachi, M.: Application of singular vector analysis to the Kuroshio large meander, J. Geophys. Res.-Oceans, 113, C07026, https://doi.org/10.1029/2007JC004476, 2008. a

Kalnay, E.: Atmospheric modeling, data assimilation and predictability, Cambridge University Press, Cambridge, ISBN 9780521796293, https://doi.org/10.1017/CBO9780511802270, 2003. a

Kerswell, R. R.: Nonlinear nonmodal stability theory, Annu. Rev. Fluid Mech., 50, 319–345, 2018. a

Lin, C.-C.: The theory of hydrodynamic stability, Cambridge University Press, Cambridge, UK, ISBN-10: 1013839366, ISBN-13: 978-1013839368, 1955. a

Liu, D. C. and Nocedal, J.: On the limited memory BFGS method for large scale optimization, Math. Program., 45, 503–528, 1989. a

Lorenz, E.: Predictability – a problem partly solved, in: Predictability of Weather and Climate, edited by: Palmer, T. and Hagedorn, R., Cambridge University Press, Cambridge, 40–58, https://doi.org/10.1017/CBO9780511617652.004, 2006. a, b, c

Lorenz, E. N.: A study of the predictability of a 28-variable atmospheric model, Tellus, 17, 321–333, 1965. a

Lorenz, E. N. and Emanuel, K. A.: Optimal sites for supplementary weather observations: Simulation with a small model, J. Atmos. Sci., 55, 399–414, 1998. a

Luo, D., Cha, J., Zhong, L., and Dai, A.: A nonlinear multiscale interaction model for atmospheric blocking: The eddy-blocking matching mechanism, Q. J. Roy. Meteor. Soc., 140, 1785–1808, 2014. a

Luo, D., Zhang, W., Zhong, L., and Dai, A.: A nonlinear theory of atmospheric blocking: A potential vorticity gradient view, J. Atmos. Sci., 76, 2399–2427, 2019. a

Mu, M.: Nonlinear singular vectors and nonlinear singular values, Sci. China Ser. D, 43, 375–385, 2000. a

Mu, M. and Wang, J.: Nonlinear fastest growing perturbation and the first kind of predictability, Sci. China Ser. D, 44, 1128–1139, 2001. a

Mu, M. and Wang, Q.: Applications of nonlinear optimization approach to atmospheric and oceanic sciences, SCIENTIA SINICA Mathematica, 47, 1207–1222, 2017. a

Mu, M., Duan, W. S., and Wang, B.: Conditional nonlinear optimal perturbation and its applications, Nonlin. Processes Geophys., 10, 493–501, https://doi.org/10.5194/npg-10-493-2003, 2003. a, b

Mu, M., Sun, L., and Dijkstra, H. A.: The sensitivity and stability of the ocean's thermohaline circulation to finite-amplitude perturbations, J. Phys. Oceanogr., 34, 2305–2315, 2004. a

Mu, M., Xu, H., and Duan, W.: A kind of initial errors related to “spring predictability barrier” for El Niño events in Zebiak–Cane model, Geophys. Res. Lett., 34, L03709, https://doi.org/10.1029/2006GL027412, 2007. a

Mu, M., Zhou, F., and Wang, H.: A method for identifying the sensitive areas in targeted observations for tropical cyclone prediction: Conditional nonlinear optimal perturbation, Mon. Weather Rev., 137, 1623–1639, 2009. a, b

Navrose, Johnson, H. G., Brion, V., Jacquin, L., and Robinet, J.-C.: Optimal perturbation for two-dimensional vortex systems: route to non-axisymmetric state, J. Fluid Mech., 855, 922–952, 2018. a

Nemirovski, A. S. and Yudin, D. B.: Problem Complexity and Method Efficiency in Optimization, translated by: Dawson, E. R., John Wiley & Sons, Chichester, New York, Brisbane, Toronto, Singapore, ISBN 0471103454, 1983. a

Ott, E., Hunt, B. R., Szunyogh, I., Zimin, A. V., Kostelich, E. J., Corazza, M., Kalnay, E., Patil, D., and Yorke, J. A.: A local ensemble Kalman filter for atmospheric data assimilation, Tellus A, 56, 415–428, 2004. a

Pazó, D., Szendro, I. G., López, J. M., and Rodríguez, M. A.: Structure of characteristic Lyapunov vectors in spatiotemporal chaos, Phys. Rev. E, 78, 016209, 2008. a

Pringle, C. C. T. and Kerswell, R. R.: Using nonlinear transient growth to construct the minimal seed for shear flow turbulence, Phys. Rev. Lett., 105, 154502, 2010. a

Qin, X. and Mu, M.: Influence of conditional nonlinear optimal perturbations sensitivity on typhoon track forecasts, Q. J. Roy. Meteor. Soc., 138, 185–197, 2012. a

Rayleigh, L.: On the stability, or instability, of certain fluid motions, P. Lond. Math. Soc., 1, 57–72, 1879. a

Samelson, R. M. and Tziperman, E.: Instability of the chaotic ENSO: The growth-phase predictability barrier, J. Atmos. Sci., 58, 3613–3625, 2001. a

Shi, B., Luo, D., and Zhang, W.: Optimal Initial Disturbance of Atmospheric Blocking: A Barotropic View, arXiv [preprint], arXiv:2210.06011, 2022. a

Sterk, A. E. and van Kekem, D. L.: Predictability of extreme waves in the Lorenz-96 model near intermittency and quasi-periodicity, Complexity, 2017, 9419024, https://doi.org/10.1155/2017/9419024, 2017. a

Sun, G. and Mu, M.: A new approach to identify the sensitivity and importance of physical parameters combination within numerical models using the Lund–Potsdam–Jena (LPJ) model as an example, Theor. Appl. Climatol., 128, 587–601, 2017. a

Thompson, C. J.: Initial conditions for optimal growth in a coupled ocean–atmosphere model of ENSO, J. Atmos. Sci., 55, 537–557, 1998. a

Trevisan, A. and Palatella, L.: On the Kalman Filter error covariance collapse into the unstable subspace, Nonlin. Processes Geophys., 18, 243–250, https://doi.org/10.5194/npg-18-243-2011, 2011. a

Vershynin, R.: High-dimensional probability: An introduction with applications in data science, vol. 47, Cambridge University Press, Cambridge, ISBN 9781108231596, https://doi.org/10.1017/9781108231596, 2018. a, b

Wang, B. and Tan, X.: Conditional nonlinear optimal perturbations: Adjoint-free calculation method and preliminary test, Mon. Weather Rev., 138, 1043–1049, 2010. a, b, c

Wang, Q. and Mu, M.: A new application of conditional nonlinear optimal perturbation approach to boundary condition uncertainty, J. Geophys. Res.-Oceans, 120, 7979–7996, 2015. a

Wang, Q., Mu, M., and Sun, G.: A useful approach to sensitivity and predictability studies in geophysical fluid dynamics: conditional non-linear optimal perturbation, Natl. Sci. Rev., 7, 214–223, 2020. a, b, c

Wu, B., Zhou, T., and Zheng, F.: EnOI-IAU initialization scheme designed for decadal climate prediction system IAP-DecPreS, J. Adv. Model. Earth Sy., 10, 342–356, 2018. a

Xue, Y., Cane, M. A., and Zebiak, S. E.: Predictability of a coupled model of ENSO using singular vector analysis. Part I: Optimal growth in seasonal background and ENSO cycles, Mon. Weather Rev., 125, 2043–2056, 1997a. a

Xue, Y., Cane, M. A., Zebiak, S. E., and Palmer, T. N.: Predictability of a coupled model of ENSO using singular vector analysis. Part II: Optimal growth and forecast skill, Mon. Weather Rev., 125, 2057–2073, 1997b. a

Yu, H., Wang, H., Meng, Z., Mu, M., Huang, X.-Y., and Zhang, X.: A WRF-based tool for forecast sensitivity to the initial perturbation: The conditional nonlinear optimal perturbations versus the first singular vector method and comparison to MM5, J. Atmos. Ocean. Tech., 34, 187–206, 2017. a

Yuan, S., Zhao, L., and Mu, B.: Parallel cooperative co-evolution based particle swarm optimization algorithm for solving conditional nonlinear optimal perturbation, in: International Conference on Neural Information Processing, Istanbul, Turkey, 10 November 2015, edited by: Arik, S., Huang, T., Lai, W., and Liu, Q., Springer, Cham, 87–95, Print-ISBN 978-3-319-26534-6, Online-ISBN 978-3-319-26535-3, https://doi.org/10.1007/978-3-319-26535-3_11, 2015. a, b

Zanna, L., Heimbach, P., Moore, A. M., and Tziperman, E.: Optimal excitation of interannual Atlantic meridional overturning circulation variability, J. Climate, 24, 413–427, 2011. a

Zebiak, S. E. and Cane, M. A.: A model El Niño–Southern Oscillation, Mon. Weather Rev., 115, 2262–2278, 1987. a

Zheng, Q., Yang, Z., Sha, J., and Yan, J.: Conditional nonlinear optimal perturbations based on the particle swarm optimization and their applications to the predictability problems, Nonlin. Processes Geophys., 24, 101–112, https://doi.org/10.5194/npg-24-101-2017, 2017.  a, b

Zu, Z., Mu, M., and Dijkstra, H. A.: Optimal initial excitations of decadal modification of the Atlantic Meridional Overturning Circulation under the prescribed heat and freshwater flux boundary conditions, J. Phys. Oceanogr., 46, 2029–2047, 2016. a

1

Generally, the statistical machine learning techniques refer to the marriage of traditional optimization methods and statistical methods, or, say, stochastic optimization methods, where the iterative behavior is governed by the distribution instead of the point due to the attention of noise. Here, the sampling algorithm used in this paper is to numerically implement the stochastic gradient descent method, which takes the sample average to obtain the inaccurate gradient.

2

Throughout the paper, the norm is specific for the energy norm. When we implement the numerical computation, or, say in the discrete case, it reduces to the Euclidean norm as v=i=1dvi2.

3

The sub-Gaussian norm of a random variable X is defined as

X ψ 2 = inf t > 0 : E exp X 2 t 2 2 .
Download
Short summary
We introduce a sample-based algorithm to obtain the conditional nonlinear optimal perturbations. Compared with the classical adjoint-based method, it is easier to implement and reduces the required storage for the basic state. When we reduce the number of samples to some extent, it reduces the computation markedly more when using the sample-based method, which can guarantee that the CNOP obtained is nearly consistent with some minor fluctuating errors oscillating in spatial distribution.