Impact of an Ensemble of Ocean Data Assimilations in ECMWF’s next generation ocean reanalysis system

Marcin Chrust
ECMWF
Shinfield Park
Reading, United Kingdom
marcin.chrust@ecmwf.int
&Anthony T. Weaver
CERFACS / CECI CNRS UMR 5318
Toulouse, France
\ANDPhilip Browne
ECMWF
Shinfield Park
Reading, United Kingdom
&Hao Zuo
ECMWF
Shinfield Park
Reading, United Kingdom
&Magdalena Alonso Balmaseda
ECMWF
Shinfield Park
Reading, United Kingdom
Abstract

An Ensemble of Data Assimilations (EDA) can provide valuable information on the analysis and short-range forecast uncertainties. The present ECMWF operational ocean analysis and reanalysis system, called ORAS5, produces an ensemble but does not exploit it for the specification of the background-error covariance matrix 𝐁𝐁\mathbf{B}bold_B, a key component of the data assimilation system. In this article, we describe EDA developments for the ocean, which take advantage of the short-range forecast ensemble for specifying, in two distinct ways, parameters of a covariance model representation of 𝐁𝐁\mathbf{B}bold_B. First, we generate a climatological ensemble over an extended period to produce seasonally varying climatological estimates of background-error variances and horizontal correlation length-scales. Second, on each assimilation cycle, we diagnose flow-dependent variances from the ensemble and blend them with the climatological estimates to form hybrid variances. We also use the ensemble to diagnose flow-dependent vertical correlation length-scales. We demonstrate for the Argo-rich period that this new, hybrid formulation of 𝐁𝐁\mathbf{B}bold_B results in a significant reduction of background errors compared to the parameterized formulation of 𝐁𝐁\mathbf{B}bold_B used in ORAS5. The new ocean EDA system will be employed in ORAS6, ECMWF’s next generation ocean reanalysis system.

Keywords hybrid data assimilation, ensemble of data assimilations, variational assimilation, background error covariances

1 Introduction

The current generation of the ECMWF ocean reanalysis system, called ORAS5 (Zuo et al., 2019), is an ensemble system, but makes no use of the ensemble of forecasts to provide climatological or flow-dependent covariance information on forecast errors. Such information could be used to improve the specification of the background-error covariance matrix 𝐁𝐁\mathbf{B}bold_B in the variational data assimilation system used to produce the ocean reanalysis. This is supported by growing evidence of the benefits of using ensembles to enrich the representation of 𝐁𝐁\mathbf{B}bold_B not only in variational data assimilation systems for the atmosphere, where it is already a very well-established practice (Buehner, 2005; Bonavita et al., 2012; Clayton et al., 2013; Berre et al., 2015; Kleist and Ide, 2015; Bonavita et al., 2016), but also for the ocean (Daget et al., 2009; Penny et al., 2015; Storto et al., 2018; Lea et al., 2022).

In this article, we describe an Ensemble of Data Assimilations (EDA) framework that will be the basis of the upcoming 6th generation of ECMWF’s ocean reanalysis system (ORAS6). While the ensemble of forecasts can be used directly to define a sample covariance matrix estimate of 𝐁𝐁\mathbf{B}bold_B (e.g., see Lorenc (2003) and Buehner (2005)), here we use it indirectly to calibrate the parameters of a modelled representation of 𝐁𝐁\mathbf{B}bold_B. To simplify the illustration of how this is achieved, we follow the standard decomposition of a covariance matrix into a correlation matrix and a diagonal matrix of standard deviations that multiplies the correlation matrix on either side. The ensemble of forecasts are used in two distinct ways. First, using the empirically parameterized representation of 𝐁𝐁\mathbf{B}bold_B from ORAS5, an 11111111-member ensemble is produced for the years 2010–2015 in order to compute seasonal climatological estimates of both standard deviations and parameters of the correlation matrix. Second, the ensemble of forecasts is also employed to estimate flow-dependent forecast errors, by updating on each assimilation cycle the background-error standard deviations and vertical correlation length-scales. We then conduct a series of data assimilation experiments for the period 2017–2021 and assess the strengths and weaknesses of the new formulation of 𝐁𝐁\mathbf{B}bold_B. We demonstrate that accounting for flow-dependent background-error covariances allows surface observations to be assimilated more effectively, including, for the first time at ECMWF, sea surface temperature (SST) observations. We show that the improvements in 𝐁𝐁\mathbf{B}bold_B coupled with the assimilation of SST have enabled us to address a long-standing issue of large temperature biases in the Gulf Stream that were present in ORAS5.

The development of the ocean EDA can also be seen as an important step towards building a coupled EDA system at ECMWF. It is expected (de Rosnay et al., 2022) that coupled data assimilation would allow for better exploitation of interface observations and for reducing the initialization imbalances of the Earth System Model. The ability of the ocean data assimilation system to exploit surface observations in an optimal way is thus of strategic importance.

The article is structured as follows. Section 2 describes methodological aspects of the new system. An overview of the covariance model used to define 𝐁𝐁\mathbf{B}bold_B is presented in Section 2.1 and is followed in Section 2.2 by a description of how the forecast ensemble is used to estimate the covariance model parameters. Section 2.3 describes how the climatologies of the covariance model parameters are constructed and subsequently combined with “errors of the day” to form a hybrid covariance model. The performance of the new system is discussed in Section 3. First, a baseline configuration, defined using an ORAS5-like 𝐁𝐁\mathbf{B}bold_B, is described in Section 3.1. The subsequent subsections illustrate the impact of using ensemble-derived parameters in 𝐁𝐁\mathbf{B}bold_B. Section 3.2 and Section 3.3 describe the impact of the hybrid variances and hybrid correlation model, respectively, while Section 3.4 describes the impact of the fully hybrid covariance model that includes both the hybrid variances and hybrid correlation model. Conclusions and outlook are given in Section 4.

2 Methodology

The new ocean EDA system is built around NEMO v4.0.6, coupled to a sea-ice model SI3 (Madec et al., 2023). The model employs a tri-polar extended ORCA grid at approximately 0.25superscript0.250.25^{\circ}0.25 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT horizontal resolution and with 75757575 vertical levels. The EDA consists of an unperturbed control member accompanied by 10101010 perturbed members. The number of perturbed members was chosen based on the consideration of computational cost and the requirement to reduce sampling noise when estimating the parameters of 𝐁𝐁\mathbf{B}bold_B. The perturbed members assimilate randomly displaced observations and are forced using hourly perturbed forcing fields derived from ERA5 (Hersbach et al., 2020). The ocean initial conditions are implicitly perturbed via the cycling process of the EDA. The model itself remains unperturbed as in ORAS5. Data assimilation is carried out using NEMOVAR (Mogensen et al., 2012) in its three-dimensional variational assimilation (3D-Var) First Guess at Appropriate Time (FGAT) configuration with a 5555-day assimilation window. The analysis increment obtained from the 3D-Var minimization is applied to the model as a tendency from the start of the window using Incremental Analysis Updates (Bloom et al., 1996). The assimilated observations contain information on temperature, salinity, sea surface height (SSH) and sea ice concentration (SIC). The details of the observing network and perturbation schemes are outside the scope of this article. The interested reader is referred to the upcoming ORAS6 article for more details (Zuo et al., 2024, in preparation). Here, we focus on the description of 𝐁𝐁\mathbf{B}bold_B and the methods for using the ensemble of forecasts to compute climatological and flow-dependent parameters for 𝐁𝐁\mathbf{B}bold_B. Details of the 𝐁𝐁\mathbf{B}bold_B formulation employed in the current operational ocean reanalysis system, ORAS5, are provided to facilitate comparison and interpretation of the results. The impact of the new 𝐁𝐁\mathbf{B}bold_B formulation is illustrated in the data-rich period 2017–2021.

2.1 The background-error covariance model

In this section, we describe the formulation of 𝐁𝐁\mathbf{B}bold_B that is used for ORAS6 and in doing so expose the covariance model parameters that are estimated using the EDA. We keep technical details to a minimum as many of them can be found in existing references (Weaver et al., 2005; Balmaseda et al., 2013; Weaver et al., 2016, 2021).

The ocean model state variables consist of potential temperature (T𝑇Titalic_T), practical salinity (S𝑆Sitalic_S), the two components of horizontal velocity (u𝑢uitalic_u and v𝑣vitalic_v), SSH (η𝜂\etaitalic_η) and SIC (a𝑎aitalic_a). In the ORAS6 3D-Var FGAT system, the control variables that are estimated through data assimilation are temperature, SIC, and the so-called unbalanced parts of salinity (Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT) and SSH (ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT). The unbalanced parts of horizontal velocity (uusubscript𝑢uu_{\rm u}italic_u start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT and vusubscript𝑣uv_{\rm u}italic_v start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT) are not considered as control variables since no information is assimilated in the 3D-Var system to correct them. The state and control vectors of the assimilation problem consist of discrete values of the state and control variables at the model grid-points.

2.1.1 Balance operator

The background-error covariance matrix is formulated in terms of a balance operator 𝐊bsubscript𝐊b\mathbf{K}_{\rm b}bold_K start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT and a covariance matrix 𝐁csubscript𝐁c\mathbf{B}_{\rm c}bold_B start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT for the control variables:

𝐁=𝐊b𝐁c𝐊bT.𝐁subscript𝐊bsubscript𝐁csuperscriptsubscript𝐊bT\mathbf{B}\,=\,\mathbf{K}_{\rm b}\,\mathbf{B}_{\rm c}\,\mathbf{K}_{\rm b}^{\rm T}.bold_B = bold_K start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT bold_B start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT bold_K start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT .

We make the fundamental assumption that the control variables are mutually uncorrelated, so that 𝐁csubscript𝐁c\mathbf{B}_{\rm c}bold_B start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT can be represented by a block-diagonal matrix

𝐁c=(𝐁TT𝟎𝟎𝟎𝟎𝐁SuSu𝟎𝟎𝟎𝟎𝐁ηuηu𝟎𝟎𝟎𝟎𝐁aa),subscript𝐁csubscript𝐁𝑇𝑇0000subscript𝐁subscript𝑆usubscript𝑆u0000subscript𝐁subscript𝜂usubscript𝜂u0000subscript𝐁𝑎𝑎\mathbf{B}_{\rm c}=\!\left(\!\!\begin{array}[]{cccc}\mathbf{B}_{TT}&\mathbf{0}% &\mathbf{0}&\mathbf{0}\\ \mathbf{0}&\mathbf{B}_{S_{\rm u}S_{\rm u}}&\mathbf{0}&\mathbf{0}\\ \mathbf{0}&\mathbf{0}&\mathbf{B}_{\eta_{\rm u}\eta_{\rm u}}&\mathbf{0}\\ \mathbf{0}&\mathbf{0}&\mathbf{0}&\mathbf{B}_{aa}\end{array}\!\!\right),bold_B start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT = ( start_ARRAY start_ROW start_CELL bold_B start_POSTSUBSCRIPT italic_T italic_T end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL end_ROW start_ROW start_CELL bold_0 end_CELL start_CELL bold_B start_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL end_ROW start_ROW start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL start_CELL bold_B start_POSTSUBSCRIPT italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL end_ROW start_ROW start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL start_CELL bold_B start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT end_CELL end_ROW end_ARRAY ) ,

where 𝐁ααsubscript𝐁𝛼𝛼\mathbf{B}_{\alpha\alpha}bold_B start_POSTSUBSCRIPT italic_α italic_α end_POSTSUBSCRIPT is a univariate covariance matrix for the control variable α=T𝛼𝑇\alpha=Titalic_α = italic_T, Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT, ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT or a𝑎aitalic_a.

Cross-covariances between the background errors of different state variables are induced by the balance operator, which has the specific structure (Weaver et al., 2005)

𝐊b=(𝐈Nhz𝟎𝟎𝟎𝐊ST𝐈Nhz𝟎𝟎𝐊ηT𝟎𝐈Nh𝟎𝟎𝟎𝟎𝐈Nh𝐊uT𝟎𝐊uηu𝟎𝐊vT𝟎𝐊vηu𝟎)subscript𝐊bsubscript𝐈subscript𝑁hz000subscript𝐊𝑆𝑇subscript𝐈subscript𝑁hz00subscript𝐊𝜂𝑇0subscript𝐈subscript𝑁h0000subscript𝐈subscript𝑁hsubscript𝐊𝑢𝑇0subscript𝐊𝑢subscript𝜂u0subscript𝐊𝑣𝑇0subscript𝐊𝑣subscript𝜂u0\mathbf{K}_{\rm b}=\!\left(\!\!\begin{array}[]{cccc}\mathbf{I}_{N_{\rm hz}}&% \mathbf{0}&\mathbf{0}&\mathbf{0}\\ \mathbf{K}_{ST}&\mathbf{I}_{N_{\rm hz}}&\mathbf{0}&\mathbf{0}\\ \mathbf{K}_{\eta T}&\mathbf{0}&\mathbf{I}_{N_{\rm h}}&\mathbf{0}\\ \mathbf{0}&\mathbf{0}&\mathbf{0}&\mathbf{I}_{N_{\rm h}}\\ \mathbf{K}_{uT}&\mathbf{0}&\mathbf{K}_{u\eta_{\rm u}}&\mathbf{0}\\ \mathbf{K}_{vT}&\mathbf{0}&\mathbf{K}_{v\eta_{\rm u}}&\mathbf{0}\end{array}\!% \!\right)bold_K start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT = ( start_ARRAY start_ROW start_CELL bold_I start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT roman_hz end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL end_ROW start_ROW start_CELL bold_K start_POSTSUBSCRIPT italic_S italic_T end_POSTSUBSCRIPT end_CELL start_CELL bold_I start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT roman_hz end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL end_ROW start_ROW start_CELL bold_K start_POSTSUBSCRIPT italic_η italic_T end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL start_CELL bold_I start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL end_ROW start_ROW start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL start_CELL bold_0 end_CELL start_CELL bold_I start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL bold_K start_POSTSUBSCRIPT italic_u italic_T end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL start_CELL bold_K start_POSTSUBSCRIPT italic_u italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL end_ROW start_ROW start_CELL bold_K start_POSTSUBSCRIPT italic_v italic_T end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL start_CELL bold_K start_POSTSUBSCRIPT italic_v italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL start_CELL bold_0 end_CELL end_ROW end_ARRAY ) (1)

where the block components 𝐊βαsubscript𝐊𝛽𝛼\mathbf{K}_{\beta\alpha}bold_K start_POSTSUBSCRIPT italic_β italic_α end_POSTSUBSCRIPT constitute the balance between the control variable α𝛼\alphaitalic_α and the state variable β𝛽\betaitalic_β, and 𝐈Nhsubscript𝐈subscript𝑁h\mathbf{I}_{N_{\rm h}}bold_I start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT end_POSTSUBSCRIPT and 𝐈Nhzsubscript𝐈subscript𝑁hz\mathbf{I}_{N_{\rm hz}}bold_I start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT roman_hz end_POSTSUBSCRIPT end_POSTSUBSCRIPT are identity matrices, Nhsubscript𝑁hN_{\rm h}italic_N start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT and Nhzsubscript𝑁hzN_{\rm hz}italic_N start_POSTSUBSCRIPT roman_hz end_POSTSUBSCRIPT corresponding to the number of two-dimensional (2D; horizontal) and three-dimensional (3D; horizontal-vertical) grid-points, respectively. The specific balance relationships (𝐊βαsubscript𝐊𝛽𝛼\mathbf{K}_{\beta\alpha}bold_K start_POSTSUBSCRIPT italic_β italic_α end_POSTSUBSCRIPT) between variables are essentially those described in Ricci et al. (2005), Weaver et al. (2005) and Balmaseda et al. (2013), but with some refinements that have been introduced over the years. The basic relations are summarized below.

The balance between T𝑇Titalic_T and S𝑆Sitalic_S (𝐊STsubscript𝐊𝑆𝑇\mathbf{K}_{ST}bold_K start_POSTSUBSCRIPT italic_S italic_T end_POSTSUBSCRIPT) involves making a local vertical adjustment to S𝑆Sitalic_S to approximately preserve the water-mass properties of the background state (Ricci et al., 2005). Changes to the water-mass properties occur through the unbalanced component Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT. The other variables (η𝜂\etaitalic_η, u𝑢uitalic_u and v𝑣vitalic_v) depend on T𝑇Titalic_T and S𝑆Sitalic_S through density, which is defined from a linear equation of state with thermal expansion and saline contraction coefficients computed from the background state. Note that only the balanced component of S𝑆Sitalic_S is used in the computation of density, which is reflected by the absence of balance between Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT and the other variables in Equation (1). The balance between η𝜂\etaitalic_η and T𝑇Titalic_T (𝐊ηTsubscript𝐊𝜂𝑇\mathbf{K}_{\eta T}bold_K start_POSTSUBSCRIPT italic_η italic_T end_POSTSUBSCRIPT) is determined by computing dynamic height at the surface relative to the ocean bottom. The unbalanced component ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT is interpreted as a barotropic contribution to SSH, that is in balance with a depth-independent geostrophic velocity component (𝐊uηusubscript𝐊𝑢subscript𝜂u\mathbf{K}_{u\eta_{\rm u}}bold_K start_POSTSUBSCRIPT italic_u italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT end_POSTSUBSCRIPT and 𝐊vηusubscript𝐊𝑣subscript𝜂u\mathbf{K}_{v\eta_{\rm u}}bold_K start_POSTSUBSCRIPT italic_v italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT end_POSTSUBSCRIPT). The remaining velocity balance (𝐊uTsubscript𝐊𝑢𝑇\mathbf{K}_{uT}bold_K start_POSTSUBSCRIPT italic_u italic_T end_POSTSUBSCRIPT and 𝐊vTsubscript𝐊𝑣𝑇\mathbf{K}_{vT}bold_K start_POSTSUBSCRIPT italic_v italic_T end_POSTSUBSCRIPT) is depth-dependent and assumed to be in geostrophic balance with a horizontal gradient of pressure computed from the hydrostatic relation. Near the Equator, the f𝑓fitalic_f-plane geostrophic balance is smoothly reduced to zero, and replaced by a β𝛽\betaitalic_β-plane geostrophic balance for the u𝑢uitalic_u-component only.

In Equation (1) the velocity balance acts as a strong constraint since there are no unbalanced velocity components in the control vector. It would be sufficient then to apply this particular balance directly to the analysis increments of T𝑇Titalic_T, Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT and ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT generated by the 3D-Var minimization. However, instead of applying the velocity balance as a post-analysis adjustment, we have formally included it in 𝐊bsubscript𝐊b\mathbf{K}_{\rm b}bold_K start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT (and hence as part of the 3D-Var minimization algorithm) in preparation for future extensions of the system that will assimilate velocity information and thus allow us to estimate the unbalanced (ageostrophic) components of velocity. There is currently no balance imposed between SIC and any of the other variables.

2.1.2 The univariate covariance model

The block covariance matrices in 𝐁csubscript𝐁c\mathbf{B}_{\rm c}bold_B start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT are factored as

𝐁αα=𝚺α𝐂αα𝚺αsubscript𝐁𝛼𝛼subscript𝚺𝛼subscript𝐂𝛼𝛼subscript𝚺𝛼\mathbf{B}_{\alpha\alpha}\,=\,\boldsymbol{\Sigma}_{\alpha}\,\mathbf{C}_{\alpha% \alpha}\,\boldsymbol{\Sigma}_{\alpha}bold_B start_POSTSUBSCRIPT italic_α italic_α end_POSTSUBSCRIPT = bold_Σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT bold_C start_POSTSUBSCRIPT italic_α italic_α end_POSTSUBSCRIPT bold_Σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT

where 𝐂ααsubscript𝐂𝛼𝛼\mathbf{C}_{\alpha\alpha}bold_C start_POSTSUBSCRIPT italic_α italic_α end_POSTSUBSCRIPT is a correlation matrix and 𝚺αsubscript𝚺𝛼\boldsymbol{\Sigma}_{\alpha}bold_Σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is a diagonal matrix of standard deviations. The standard deviations for all control variables except for SIC are estimated using the EDA as described in the next section. Each correlation matrix is modelled using a diffusion operator, formed by discretising the pseudo-time derivative of a diffusion equation with an implicit (Euler backward) scheme. In what follows, we refer to this as an implicit diffusion operator. The total number of diffusion steps M𝑀Mitalic_M primarily controls the spectral decay rate of the smoothing kernel of the diffusion operator at high wavenumbers. These smoothing kernels are covariance functions from the Matérn family (Guttorp and Gneiting, 2006; Weaver and Mirouze, 2013). Here, we have set M𝑀Mitalic_M to a value of 10 (for all variables) so that the spatial correlations are approximately Gaussian (see Figure A1 in Goux et al. (2024)).

The correlation matrix has the symmetric, factored form

𝐂αα=𝚪α𝐕α𝐖α1𝐕αT𝚪αsubscript𝐂𝛼𝛼subscript𝚪𝛼subscript𝐕𝛼superscriptsubscript𝐖𝛼1superscriptsubscript𝐕𝛼Tsubscript𝚪𝛼\mathbf{C}_{\alpha\alpha}\,=\,\boldsymbol{\Gamma}_{\alpha}\,\mathbf{V}_{\alpha% }\,\mathbf{W}_{\alpha}^{-1}\,\mathbf{V}_{\alpha}^{\rm T}\,\boldsymbol{\Gamma}_% {\alpha}bold_C start_POSTSUBSCRIPT italic_α italic_α end_POSTSUBSCRIPT = bold_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT bold_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT bold_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT (2)

where 𝐕αsubscript𝐕𝛼\mathbf{V}_{\alpha}bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is a discrete representation of a diffusion operator as detailed below; 𝚪αsubscript𝚪𝛼\boldsymbol{\Gamma}_{\alpha}bold_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is a diagonal matrix containing normalization factors to ensure that the diagonal elements of 𝐂ααsubscript𝐂𝛼𝛼\mathbf{C}_{\alpha\alpha}bold_C start_POSTSUBSCRIPT italic_α italic_α end_POSTSUBSCRIPT are all approximately equal to one; and 𝐖αsubscript𝐖𝛼\mathbf{W}_{\alpha}bold_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is a diagonal matrix of either area elements (𝐖hsubscript𝐖h\mathbf{W}_{\rm h}bold_W start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT) or volume elements (𝐖hzsubscript𝐖hz\mathbf{W}_{\rm hz}bold_W start_POSTSUBSCRIPT roman_hz end_POSTSUBSCRIPT) depending on the variable:

𝐖α={𝐖hifα=ηuora,𝐖hzifα=TorSu.subscript𝐖𝛼casessubscript𝐖hif𝛼subscript𝜂uor𝑎subscript𝐖hzif𝛼𝑇orsubscript𝑆u\mathbf{W}_{\alpha}=\left\{\begin{array}[]{lcl}\mathbf{W}_{\rm h}&\mbox{if}&% \alpha=\eta_{\rm u}\;\mbox{or}\;a,\\ \mathbf{W}_{\rm hz}&\mbox{if}&\alpha=T\;\mbox{or}\;S_{\rm u}.\end{array}\right.bold_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = { start_ARRAY start_ROW start_CELL bold_W start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT end_CELL start_CELL if end_CELL start_CELL italic_α = italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT or italic_a , end_CELL end_ROW start_ROW start_CELL bold_W start_POSTSUBSCRIPT roman_hz end_POSTSUBSCRIPT end_CELL start_CELL if end_CELL start_CELL italic_α = italic_T or italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT . end_CELL end_ROW end_ARRAY

The combined matrix operator 𝐕α𝐖α1𝐕αTsubscript𝐕𝛼superscriptsubscript𝐖𝛼1superscriptsubscript𝐕𝛼T\mathbf{V}_{\alpha}\mathbf{W}_{\alpha}^{-1}\mathbf{V}_{\alpha}^{\rm T}bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT bold_W start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT represents the full M𝑀Mitalic_M-step diffusion process, while the components 𝐕αsubscript𝐕𝛼\mathbf{V}_{\alpha}bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT and 𝐕αTsuperscriptsubscript𝐕𝛼T\mathbf{V}_{\alpha}^{\rm T}bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT constitute diffusion over half the number of steps. Horizontal and vertical correlations are modelled using a 2D and one-dimensional (1D) implicit diffusion operator, respectively. Horizontal-vertical (3D) correlations are modelled by combining the horizontal and vertical implicit diffusion operators. Within 𝐕αsubscript𝐕𝛼\mathbf{V}_{\alpha}bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, the combined operator, which is used for T𝑇Titalic_T and Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT, is formed by first applying one step of the 1D implicit diffusion operator at each horizontal grid-point (an operation denoted by 𝐅α,zsubscript𝐅𝛼z\mathbf{F}_{\alpha,{\rm z}}bold_F start_POSTSUBSCRIPT italic_α , roman_z end_POSTSUBSCRIPT), followed by applying one step of the 2D implicit diffusion operator in each vertical level (an operation denoted by 𝐅α,hsubscript𝐅𝛼h\mathbf{F}_{\alpha,{\rm h}}bold_F start_POSTSUBSCRIPT italic_α , roman_h end_POSTSUBSCRIPT). We refer to this 3D diffusion formulation as 2D×\times×1D. For ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT and a𝑎aitalic_a, only the horizontal diffusion operator is needed. In terms of the one-step components, 𝐕αsubscript𝐕𝛼\mathbf{V}_{\alpha}bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT can be expressed as

𝐕α={𝐅α,hM/2ifα=ηuora,(𝐅α,h𝐅α,z)M/2ifα=TorSu.subscript𝐕𝛼casessuperscriptsubscript𝐅𝛼h𝑀2if𝛼subscript𝜂uor𝑎superscriptsubscript𝐅𝛼hsubscript𝐅𝛼z𝑀2if𝛼𝑇orsubscript𝑆u\mathbf{V}_{\alpha}=\left\{\begin{array}[]{lcl}\mathbf{F}_{\alpha,{\rm h}}^{M/% 2}&\mbox{if}&\alpha=\eta_{\rm u}\;\mbox{or}\;a,\\ \left(\mathbf{F}_{\alpha,{\rm h}}\,\mathbf{F}_{\alpha,{\rm z}}\right)^{M/2}&% \mbox{if}&\alpha=T\;\mbox{or}\;S_{\rm u}.\end{array}\right.bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = { start_ARRAY start_ROW start_CELL bold_F start_POSTSUBSCRIPT italic_α , roman_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M / 2 end_POSTSUPERSCRIPT end_CELL start_CELL if end_CELL start_CELL italic_α = italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT or italic_a , end_CELL end_ROW start_ROW start_CELL ( bold_F start_POSTSUBSCRIPT italic_α , roman_h end_POSTSUBSCRIPT bold_F start_POSTSUBSCRIPT italic_α , roman_z end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_M / 2 end_POSTSUPERSCRIPT end_CELL start_CELL if end_CELL start_CELL italic_α = italic_T or italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT . end_CELL end_ROW end_ARRAY (3)

Rather than constructing the 3D correlation operator as a product of 2D and 1D diffusion operators as in Equation (3), an alternative and arguably more natural approach is to construct it directly from a discretised 3D implicit diffusion equation (see Section 2.2.3 in Weaver et al. (2021)). The resulting correlation model has similar smoothness properties to Equation (3) (see Figure 18 in Weaver et al. (2021)) but is computationally more expensive and therefore has not been used for ORAS6.

Consider the application of the 2D×\times×1D implicit diffusion operator to a vector 𝜶m1subscript𝜶𝑚1\boldsymbol{\alpha}_{m-1}bold_italic_α start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT where the subscript m𝑚mitalic_m denotes the step counter in the diffusion process. Performing one step of 2D implicit diffusion involves solving a linear system

𝐀hk𝜶mhk=𝜶m1hksubscript𝐀subscripth𝑘superscriptsubscript𝜶𝑚subscripth𝑘superscriptsubscript𝜶𝑚1subscripth𝑘\mathbf{A}_{{\rm h}_{k}}\boldsymbol{\alpha}_{m}^{{\rm h}_{k}}=\boldsymbol{% \alpha}_{m-1}^{{\rm h}_{k}}bold_A start_POSTSUBSCRIPT roman_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT bold_italic_α start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT = bold_italic_α start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT (4)

where 𝜶m1hksuperscriptsubscript𝜶𝑚1subscripth𝑘\boldsymbol{\alpha}_{m-1}^{{\rm h}_{k}}bold_italic_α start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUPERSCRIPT is the sub-vector of 𝜶m1subscript𝜶𝑚1\boldsymbol{\alpha}_{m-1}bold_italic_α start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT associated with model level k𝑘kitalic_k. Similarly, performing one step of 1D implicit diffusion involves solving, at each horizontal grid-point (i,j)𝑖𝑗(i,j)( italic_i , italic_j ), a linear system

𝐀zij𝜶mzij=𝜶m1zijsubscript𝐀subscriptz𝑖𝑗superscriptsubscript𝜶𝑚subscriptz𝑖𝑗superscriptsubscript𝜶𝑚1subscriptz𝑖𝑗\mathbf{A}_{{\rm z}_{ij}}\boldsymbol{\alpha}_{m}^{{\rm z}_{ij}}=\boldsymbol{% \alpha}_{m-1}^{{\rm z}_{ij}}bold_A start_POSTSUBSCRIPT roman_z start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT bold_italic_α start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_z start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT = bold_italic_α start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_z start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT (5)

where 𝜶m1zijsuperscriptsubscript𝜶𝑚1subscriptz𝑖𝑗\boldsymbol{\alpha}_{m-1}^{{\rm z}_{ij}}bold_italic_α start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_z start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT is the sub-vector of 𝜶m1subscript𝜶𝑚1\boldsymbol{\alpha}_{m-1}bold_italic_α start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT associated with the vertical column at (i,j)𝑖𝑗(i,j)( italic_i , italic_j ). The matrices 𝐀hksubscript𝐀subscripth𝑘\mathbf{A}_{{\rm h}_{k}}bold_A start_POSTSUBSCRIPT roman_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT and 𝐀zijsubscript𝐀subscriptz𝑖𝑗\mathbf{A}_{{\rm z}_{ij}}bold_A start_POSTSUBSCRIPT roman_z start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT are positive definite and self-adjoint with respect to the inner product 𝜶iT𝐖hz𝜶jsuperscriptsubscript𝜶𝑖Tsubscript𝐖hzsubscript𝜶𝑗\boldsymbol{\alpha}_{i}^{\rm T}\mathbf{W}_{\rm hz}\boldsymbol{\alpha}_{j}bold_italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT bold_W start_POSTSUBSCRIPT roman_hz end_POSTSUBSCRIPT bold_italic_α start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT.

The matrices 𝐀hksubscript𝐀subscripth𝑘\mathbf{A}_{{\rm h}_{k}}bold_A start_POSTSUBSCRIPT roman_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT and 𝐀zijsubscript𝐀subscriptz𝑖𝑗\mathbf{A}_{{\rm z}_{ij}}bold_A start_POSTSUBSCRIPT roman_z start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT are constructed from the horizontal and vertical components of the 3D elliptic operator associated with the implicitly discretised 3D diffusion equation. In NEMOVAR, the analysis is performed directly on the ORCA grid, and the spatial discretisation techniques used for the diffusion equation are inherited from NEMO (Madec et al., 2023). Adopting NEMO notation, the continuous counterparts of 𝐀hksubscript𝐀subscripth𝑘\mathbf{A}_{{\rm h}_{k}}bold_A start_POSTSUBSCRIPT roman_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT and 𝐀zijsubscript𝐀subscriptz𝑖𝑗\mathbf{A}_{{\rm z}_{ij}}bold_A start_POSTSUBSCRIPT roman_z start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT are the elliptic operators

Ah:=I1e1e2e3(i(κ11e2e3e1i)+j(κ22e1e3e2j))assignsubscript𝐴h𝐼1subscript𝑒1subscript𝑒2subscript𝑒3𝑖subscript𝜅11subscript𝑒2subscript𝑒3subscript𝑒1𝑖𝑗subscript𝜅22subscript𝑒1subscript𝑒3subscript𝑒2𝑗A_{{\rm h}}:=I-\frac{1}{e_{1}e_{2}e_{3}}\left(\frac{\partial}{\partial i}\left% (\kappa_{11}\frac{e_{2}e_{3}}{e_{1}}\frac{\partial}{\partial i}\right)+\frac{% \partial}{\partial j}\left(\kappa_{22}\frac{e_{1}e_{3}}{e_{2}}\frac{\partial}{% \partial j}\right)\right)italic_A start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT := italic_I - divide start_ARG 1 end_ARG start_ARG italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG ( divide start_ARG ∂ end_ARG start_ARG ∂ italic_i end_ARG ( italic_κ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT divide start_ARG italic_e start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG start_ARG italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ end_ARG start_ARG ∂ italic_i end_ARG ) + divide start_ARG ∂ end_ARG start_ARG ∂ italic_j end_ARG ( italic_κ start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT divide start_ARG italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG start_ARG italic_e start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ end_ARG start_ARG ∂ italic_j end_ARG ) ) (6)

and

Az:=I1e1e2e3(k(κ33e1e2e3k))assignsubscript𝐴z𝐼1subscript𝑒1subscript𝑒2subscript𝑒3𝑘subscript𝜅33subscript𝑒1subscript𝑒2subscript𝑒3𝑘A_{{\rm z}}:=I-\frac{1}{e_{1}e_{2}e_{3}}\left(\frac{\partial}{\partial k}\left% (\kappa_{33}\frac{e_{1}e_{2}}{e_{3}}\frac{\partial}{\partial k}\right)\right)italic_A start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT := italic_I - divide start_ARG 1 end_ARG start_ARG italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG ( divide start_ARG ∂ end_ARG start_ARG ∂ italic_k end_ARG ( italic_κ start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT divide start_ARG italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_e start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ end_ARG start_ARG ∂ italic_k end_ARG ) ) (7)

where I𝐼Iitalic_I is the identity operator, and e1subscript𝑒1e_{1}italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, e2subscript𝑒2e_{2}italic_e start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and e3subscript𝑒3e_{3}italic_e start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT are the scale factors that define the local deformation of the general orthogonal curvilinear coordinates (i,j,k)𝑖𝑗𝑘(i,j,k)( italic_i , italic_j , italic_k ) in the i𝑖iitalic_i, j𝑗jitalic_j and k𝑘kitalic_k direction, respectively. Note that the factor e1e2e3subscript𝑒1subscript𝑒2subscript𝑒3e_{1}e_{2}e_{3}italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT corresponds to a volume element and its value at each grid-point constitutes a diagonal element of 𝐖hzsubscript𝐖hz\mathbf{W}_{\rm hz}bold_W start_POSTSUBSCRIPT roman_hz end_POSTSUBSCRIPT. The elliptic operator associated with the 2D diffusion-based correlation operators for ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT and a𝑎aitalic_a is the same as Equation (6) but with the scale factor e3subscript𝑒3e_{3}italic_e start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT omitted. The diffusion coefficients κ11subscript𝜅11\kappa_{11}italic_κ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT and κ22subscript𝜅22\kappa_{22}italic_κ start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT in Equation (6) are the parameters that control the length-scales of the horizontal correlations in the i𝑖iitalic_i and j𝑗jitalic_j directions, while the diffusion coefficient κ33subscript𝜅33\kappa_{33}italic_κ start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT is the parameter that controls the length-scale of the vertical correlations. These parameters depend on the three spatial coordinates (i,j,k)𝑖𝑗𝑘(i,j,k)( italic_i , italic_j , italic_k ). More generally, these coefficients constitute the diagonal elements of a 3D diffusion tensor at each grid-point. In the next section, we describe how these parameters are estimated from the EDA.

Equation (5) is solved exactly at each grid-point (i,j)𝑖𝑗(i,j)( italic_i , italic_j ) by decomposing 𝐀zijsubscript𝐀subscriptz𝑖𝑗\mathbf{A}_{{\rm z}_{ij}}bold_A start_POSTSUBSCRIPT roman_z start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT into Cholesky factors and using forward and backward substitution. Solving Equation (4) is the most computationally demanding operation in 𝐁𝐁\mathbf{B}bold_B. In NEMOVAR, it is solved approximately in each level k𝑘kitalic_k using a linear solver based on the Chebyshev Iteration (Weaver et al., 2016). The same, fixed number of iterations are used in 𝐕αsubscript𝐕𝛼\mathbf{V}_{\alpha}bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT and 𝐕αTsuperscriptsubscript𝐕𝛼T\mathbf{V}_{\alpha}^{\rm T}bold_V start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT in order to preserve the symmetric property of 𝐂ααsubscript𝐂𝛼𝛼\mathbf{C}_{\alpha\alpha}bold_C start_POSTSUBSCRIPT italic_α italic_α end_POSTSUBSCRIPT to within machine precision. The number of iterations for each diffusion step is pre-computed by solving the 2D implicit diffusion equation with a random right-hand side and diagnosing the iterations required to achieve a three orders of magnitude reduction of the 2-norm of the residual relative to the 2-norm of the right-hand side. The total number of iterations ranges between 8 and 24 for T𝑇Titalic_T and Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT, and 25 and 36 for ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT and a𝑎aitalic_a.

2.2 Estimation of variances and correlation model parameters

The ensemble perturbations from the EDA are used to calibrate parameters of the background-error covariance models for T𝑇Titalic_T, Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT and ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT. These parameters consist of the standard deviations σ𝜎\sigmaitalic_σ and the coefficients κ11subscript𝜅11\kappa_{11}italic_κ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT, κ22subscript𝜅22\kappa_{22}italic_κ start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT and κ33subscript𝜅33\kappa_{33}italic_κ start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT of the diffusion-based correlation model, at each grid-point. The total number of discrete parameters to estimate is 2 variables ×\times× 4 parameters ×\times× Nhzsubscript𝑁hzN_{\rm hz}italic_N start_POSTSUBSCRIPT roman_hz end_POSTSUBSCRIPT grid-points +++ 1 variable ×\times× 3 parameters ×\times× Nhsubscript𝑁hN_{\rm h}italic_N start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT grid-points, where Nhzsubscript𝑁hzN_{\rm hz}italic_N start_POSTSUBSCRIPT roman_hz end_POSTSUBSCRIPT and Nhsubscript𝑁hN_{\rm h}italic_N start_POSTSUBSCRIPT roman_h end_POSTSUBSCRIPT are understood here to include only the active (ocean) points. In the ORAS6 configuration, this amounts to approximately 4.3×1084.3superscript1084.3\times 10^{8}4.3 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT discrete parameters.

The inverse of the balance operator is required to compute perturbations of the control variables from the perturbations of the state variables provided by the ensemble. In particular, we need to compute the perturbations of T𝑇Titalic_T, Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT and ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT by removing the balanced component from the perturbations of T𝑇Titalic_T, S𝑆Sitalic_S and η𝜂\etaitalic_η. This computation requires applying the inverse of the upper 3×\times×3 block matrix in Equation (1), which is straightforward given its lower triangular structure. Let 𝐊^bsubscript^𝐊b\widehat{\mathbf{K}}_{\rm b}over^ start_ARG bold_K end_ARG start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT denote this block sub-matrix and let {𝐱p}subscript𝐱𝑝\left\{\mathbf{x}_{p}\right\}{ bold_x start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT }, p=0,,Ne𝑝0subscript𝑁ep=0,\ldots,N_{\rm e}italic_p = 0 , … , italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT, denote the ensemble of background states of T𝑇Titalic_T, S𝑆Sitalic_S and η𝜂\etaitalic_η where p=0𝑝0p=0italic_p = 0 corresponds to the unperturbed member. If 𝐞psuperscriptsubscript𝐞𝑝\mathbf{e}_{p}^{\prime}bold_e start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT is the vector containing the perturbations of T𝑇Titalic_T, S𝑆Sitalic_S and η𝜂\etaitalic_η, centred about their mean,

𝐞p=𝐱p1Nel=1Ne𝐱l,superscriptsubscript𝐞𝑝subscript𝐱𝑝1subscript𝑁esuperscriptsubscript𝑙1subscript𝑁esubscript𝐱𝑙\mathbf{e}_{p}^{\prime}=\mathbf{x}_{p}-\frac{1}{N_{\rm e}}\sum_{l=1}^{N_{\rm e% }}\mathbf{x}_{l},bold_e start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = bold_x start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT end_POSTSUPERSCRIPT bold_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ,

then the centred control vector perturbations ϵpsuperscriptsubscriptbold-italic-ϵ𝑝\boldsymbol{\epsilon}_{p}^{\prime}bold_italic_ϵ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT are defined as

ϵp=𝐊^b1𝐞p,p=1,,Ne.formulae-sequencesuperscriptsubscriptbold-italic-ϵ𝑝superscriptsubscript^𝐊b1superscriptsubscript𝐞𝑝𝑝1subscript𝑁e\boldsymbol{\epsilon}_{p}^{\prime}=\widehat{\mathbf{K}}_{\rm b}^{-1}\mathbf{e}% _{p}^{\prime},\hskip 28.45274ptp=1,\ldots,N_{\rm e}.bold_italic_ϵ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = over^ start_ARG bold_K end_ARG start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_e start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_p = 1 , … , italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT .

The unperturbed member 𝐱0subscript𝐱0\mathbf{x}_{0}bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is used as the reference state for the T𝑇Titalic_T-S𝑆Sitalic_S and density relationships in 𝐊^bsubscript^𝐊b\widehat{\mathbf{K}}_{\rm b}over^ start_ARG bold_K end_ARG start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT and therefore is excluded from the ensemble average.

A sample estimate of the variances of T𝑇Titalic_T, Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT and ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT is computed on each cycle from the Ne=10subscript𝑁e10N_{\rm e}=10italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT = 10 ensemble perturbations ϵpsuperscriptsubscriptbold-italic-ϵ𝑝\boldsymbol{\epsilon}_{p}^{\prime}bold_italic_ϵ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. To reduce sampling error, the variances are filtered in each level using a 2D implicit diffusion operator with a horizontal filtering length-scale determined using the method based on optimality criteria for linear filtering of covariances derived by Ménétrier et al. (2015). The procedure for computing a sample estimate of the diffusion coefficients is more complicated and is detailed in Section 3 of Weaver et al. (2021). Only the main points are summarized here.

Central to the method is the relationship between Matérn correlation functions and the smoothing kernels of the implicit diffusion operator. The local curvature of the correlation function near its peak can be used to characterize the spatial range of the function. For an anisotropic function, the local curvature is defined by the Local Correlation (Hessian) Tensor (LCT). If 𝑯𝑯\boldsymbol{H}bold_italic_H denotes the LCT of an at least twice differentiable correlation function c(r)𝑐𝑟c(r)italic_c ( italic_r ) then 𝑯:=Tc|r=0assign𝑯evaluated-atsuperscriptT𝑐𝑟0\boldsymbol{H}:=-\nabla\nabla^{\rm T}c|_{r=0}bold_italic_H := - ∇ ∇ start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT italic_c | start_POSTSUBSCRIPT italic_r = 0 end_POSTSUBSCRIPT where r𝑟ritalic_r is Euclidean distance. For Matérn functions in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, we have the relation (Weaver and Mirouze, 2013)

𝜿=(12M4)𝑯1𝜿12𝑀4superscript𝑯1\boldsymbol{\kappa}\,=\,\left(\frac{1}{2M-4}\right)\boldsymbol{H}^{-1}bold_italic_κ = ( divide start_ARG 1 end_ARG start_ARG 2 italic_M - 4 end_ARG ) bold_italic_H start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (8)

where 𝜿𝜿\boldsymbol{\kappa}bold_italic_κ is the diffusion tensor of a 2D implicit diffusion operator. The LCT inverse, 𝑯1superscript𝑯1\boldsymbol{H}^{-1}bold_italic_H start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, for anisotropic functions can be interpreted as the tensor generalisation of the Daley length-scale for isotropic correlation functions (Daley, 1991).

In 2D, 𝑯𝑯\boldsymbol{H}bold_italic_H is a 2×2222\times 22 × 2 symmetric, positive-definite matrix with elements Hpqsubscript𝐻𝑝𝑞H_{pq}italic_H start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT, p,q=1,2formulae-sequence𝑝𝑞12p,q=1,2italic_p , italic_q = 1 , 2. This matrix can be approximated at each grid point from a sample estimate of the tensor product of the local gradient of the ensemble perturbations normalized by their sample standard deviation (Michel et al., 2016):

𝑯𝑯~=1Ne1p=1Ne(ϵp/σ~)((ϵp/σ~))T\boldsymbol{H}\,\approx\,\widetilde{\boldsymbol{H}}=\frac{1}{N_{\rm e}-1}\sum_% {p=1}^{N_{\rm e}}\nabla\big{(}\epsilon^{\prime}_{p}/\widetilde{\sigma}\big{)}% \,\big{(}\nabla\big{(}\epsilon^{\prime}_{p}/\widetilde{\sigma}\big{)}\,\big{)}% ^{\rm T}bold_italic_H ≈ over~ start_ARG bold_italic_H end_ARG = divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT - 1 end_ARG ∑ start_POSTSUBSCRIPT italic_p = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ∇ ( italic_ϵ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / over~ start_ARG italic_σ end_ARG ) ( ∇ ( italic_ϵ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / over~ start_ARG italic_σ end_ARG ) ) start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT (9)

where

σ~=1Ne1p=1Neϵp2.~𝜎1subscript𝑁e1superscriptsubscript𝑝1subscript𝑁esuperscriptsubscriptsuperscriptitalic-ϵ𝑝2\widetilde{\sigma}\,=\,\sqrt{\frac{1}{N_{\rm e}-1}\sum_{p=1}^{N_{\rm e}}{% \epsilon^{\prime}_{p}}^{2}}.over~ start_ARG italic_σ end_ARG = square-root start_ARG divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT - 1 end_ARG ∑ start_POSTSUBSCRIPT italic_p = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_ϵ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG .

The derivatives in Equation (9) are estimated numerically using centred finite-differences as described in Weaver et al. (2021). As with the variances, each of the estimated LCT elements are filtered in each model level using a 2D diffusion operator with a constant horizontal filtering length-scale, where the procedure of Michel et al. (2016) has been adopted in order to preserve positive definiteness of the LCT. After filtering, the LCT is inverted at each grid-point to obtain the Daley tensor and hence an estimate of the diffusion tensor from Equation (8). As evident from Equation (6), the current 2D implicit diffusion operator does not account for the cross-derivative terms that are needed to exploit the non-diagonal elements of the tensor. To compensate, the diagonal elements are re-scaled such that the determinant of the adjusted diagonal tensor equals the determinant of the estimated non-diagonal tensor:

κ11=κ~111κ~122/κ~11κ~22κ22=κ~221κ~122/κ~11κ~22}casessubscript𝜅11subscript~𝜅111superscriptsubscript~𝜅122subscript~𝜅11subscript~𝜅22subscript𝜅22subscript~𝜅221superscriptsubscript~𝜅122subscript~𝜅11subscript~𝜅22\displaystyle\left.\begin{array}[]{ccl}\kappa_{11}&\!\!=\!\!&\widetilde{\kappa% }_{11}\,\sqrt{1-\widetilde{\kappa}_{12}^{2}/\widetilde{\kappa}_{11}\widetilde{% \kappa}_{22}}\vspace{3mm}\\ \kappa_{22}&\!\!=\!\!&\widetilde{\kappa}_{22}\,\sqrt{1-\widetilde{\kappa}_{12}% ^{2}/\widetilde{\kappa}_{11}\widetilde{\kappa}_{22}}\end{array}\right\}start_ARRAY start_ROW start_CELL italic_κ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT end_CELL start_CELL = end_CELL start_CELL over~ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT square-root start_ARG 1 - over~ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / over~ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT over~ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL italic_κ start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT end_CELL start_CELL = end_CELL start_CELL over~ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT square-root start_ARG 1 - over~ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / over~ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT over~ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT end_ARG end_CELL end_ROW end_ARRAY } (12)

where the diffusion coefficients with the tilde are the original estimates. The final values of κ11subscript𝜅11\kappa_{11}italic_κ start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT and κ22subscript𝜅22\kappa_{22}italic_κ start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT are obtained after applying procedures to bound their minimum-allowed values and to reduce their values in semi-enclosed seas and close to boundaries where the filtered sample estimates could be excessively large (Weaver et al., 2021).

A similar estimation and filtering procedure is used to determine the vertical diffusion coefficients for the 1D diffusion operator (Equation (7)). Instead of Equation (8), we make use of the relation

κ33=(12M3)H331,subscript𝜅3312𝑀3superscriptsubscript𝐻331\kappa_{33}\,=\,\left(\frac{1}{2M-3}\right){H}_{33}^{-1},italic_κ start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT = ( divide start_ARG 1 end_ARG start_ARG 2 italic_M - 3 end_ARG ) italic_H start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ,

which has a slightly modified scaling factor that is appropriate for Matérn functions in \mathbb{R}blackboard_R. The element H331superscriptsubscript𝐻331H_{33}^{-1}italic_H start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT can be interpreted as the square of the vertical (Daley) length-scale. It is estimated at each grid-point from the inverse of a sample estimate of the square of the vertical derivative of the ensemble perturbations normalized by their sample standard deviation:

H33H~33=1Ne1p=1Ne(1e3(ϵp/σ~)k)2.subscript𝐻33subscript~𝐻331subscript𝑁e1superscriptsubscript𝑝1subscript𝑁esuperscript1subscript𝑒3subscriptsuperscriptitalic-ϵ𝑝~𝜎𝑘2H_{33}\,\approx\,\widetilde{H}_{33}=\frac{1}{N_{\rm e}-1}\sum_{p=1}^{N_{\rm e}% }\left(\frac{1}{e_{3}}\frac{\partial(\epsilon^{\prime}_{p}/\widetilde{\sigma})% }{\partial k}\right)^{\!2}.italic_H start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT ≈ over~ start_ARG italic_H end_ARG start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT - 1 end_ARG ∑ start_POSTSUBSCRIPT italic_p = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( divide start_ARG 1 end_ARG start_ARG italic_e start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ ( italic_ϵ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / over~ start_ARG italic_σ end_ARG ) end_ARG start_ARG ∂ italic_k end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT .

Before inverting H~33subscript~𝐻33\widetilde{H}_{33}over~ start_ARG italic_H end_ARG start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT, the estimate is filtered using a 1D diffusion operator equipped with an optimally-estimated vertical filtering length-scale (Ménétrier et al., 2015). Finally, κ33subscript𝜅33\kappa_{33}italic_κ start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT is adjusted to bound the minimum-allowed value, and to avoid excessively large values in the mixed layer and in shallow-sea regions, as explained in Weaver et al. (2021).

2.3 Hybrid 𝐁𝐁\mathbf{B}bold_B parameters

In order to produce more robust estimates of the covariance model parameters, we have developed a procedure to hybridize the flow-dependent ensemble estimates with climatological and parameterized values. Hybrid variances are built by combining climatological, flow-dependent and parameterized variances, while the hybridization of the diffusion coefficients is defined somewhat differently. Here, it involves using climatological horizontal diffusion coefficients in the 2D diffusion operator and flow-dependent vertical diffusion coefficients in the 1D diffusion operator. In what follows, we will loosely refer to this way of combining climatological and flow-dependent diffusion coefficients as the hybrid diffusion tensor formulation. Climatological statistics were computed over the 6–year period 2010–2015 for each season using an 11-member (10 perturbed + 1 unperturbed) ensemble system equipped with the previous generation, parameterized 𝐁𝐁\mathbf{B}bold_B (i.e. from ORAS5). The statistics were obtained by aggregating the filtered variances and filtered LCT elements across all the assimilation cycles in a season across the considered years. On average, the total effective ensemble size to compute the seasonal climatology was 1080 (10 perturbed member ×\times× 18 cycles per season ×\times× 6 years). Finally, the climatological diffusion coefficients are adjusted using the same procedures described in Section 2.2 for the cycle-dependent diffusion coefficients.

2.3.1 Hybrid variances

The background-error variances for T𝑇Titalic_T are parameterized in ORAS5 (Zuo et al., 2015) in terms of the vertical derivative of the background temperature field in order to concentrate the largest variances at the level of the thermocline. The background-error variances for Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT are parameterized using the background mixed-layer depth in order to focus large variances in the mixed layer where the T-S balance relationship is not applied. The background-error variances for ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT are parameterized using a latitude dependent function to account for the importance of the barotropic component in the extra-tropics.

As noted earlier, the employed ensemble generation scheme still relies on the deterministic model and the only sources of uncertainty that it reflects are those attributed to representativeness error of observations and the uncertainty in surface forcing. For that reason, and also owing to the eddy-permitting horizontal resolution of the model not being sufficient to resolve ocean variability at the eddy scales and below, relying on ensemble statistics alone for the specification of the background-error variances is not possible. The ensemble spread outside of eddy-active regions, like the Western Boundary Current (WBC) or the Antarctic Circumpolar Current (ACC) regions, is very small and not representative of diagnosed background forecast errors. To mitigate this issue, we combine the seasonal climatological standard deviations with the ORAS5 parameterized standard deviations using a smooth maximum function:

σm=1hlog(ehwcσc+ehwpσp1)subscript𝜎m1superscript𝑒subscript𝑤csubscript𝜎csuperscript𝑒subscript𝑤psubscript𝜎p1\sigma_{\textrm{m}}=\frac{1}{h}\log\left(e^{hw_{\textrm{c}}\sigma_{\textrm{c}}% }+e^{hw_{\textrm{p}}\sigma_{\textrm{p}}}-1\right)italic_σ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_h end_ARG roman_log ( italic_e start_POSTSUPERSCRIPT italic_h italic_w start_POSTSUBSCRIPT c end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT italic_h italic_w start_POSTSUBSCRIPT p end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT p end_POSTSUBSCRIPT end_POSTSUPERSCRIPT - 1 ) (13)

where wcsubscript𝑤cw_{\textrm{c}}italic_w start_POSTSUBSCRIPT c end_POSTSUBSCRIPT and wpsubscript𝑤pw_{\textrm{p}}italic_w start_POSTSUBSCRIPT p end_POSTSUBSCRIPT are dimensionless weighting factors for the climatological and parameterized standard deviations, σcsubscript𝜎c\sigma_{\textrm{c}}italic_σ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT and σpsubscript𝜎p\sigma_{\textrm{p}}italic_σ start_POSTSUBSCRIPT p end_POSTSUBSCRIPT, respectively, and hhitalic_h is a hardness factor, which has the property that as hh\rightarrow\inftyitalic_h → ∞, σmsubscript𝜎m\sigma_{\textrm{m}}italic_σ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT approaches the hard maximum, max(wcσc,wpσp)subscript𝑤csubscript𝜎csubscript𝑤psubscript𝜎p\max\big{(}w_{\textrm{c}}\sigma_{\textrm{c}},w_{\textrm{p}}\sigma_{\textrm{p}}% \big{)}roman_max ( italic_w start_POSTSUBSCRIPT c end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT p end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT p end_POSTSUBSCRIPT ). The hardness factor has physical units inverse to those of the standard deviation. We set h=1010h=10italic_h = 10 K-1 for T𝑇Titalic_T and h=1010h=10italic_h = 10 psu-1 for Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT, and h=100100h=100italic_h = 100 m-1 for ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT. The factor 1 in the argument of the logarithmic function ensures that σm=wcσcsubscript𝜎msubscript𝑤csubscript𝜎c\sigma_{\textrm{m}}=w_{\textrm{c}}\sigma_{\textrm{c}}italic_σ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT = italic_w start_POSTSUBSCRIPT c end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT when σp=0subscript𝜎p0\sigma_{\textrm{p}}=0italic_σ start_POSTSUBSCRIPT p end_POSTSUBSCRIPT = 0 and σm=wpσpsubscript𝜎msubscript𝑤psubscript𝜎p\sigma_{\textrm{m}}=w_{\textrm{p}}\sigma_{\textrm{p}}italic_σ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT = italic_w start_POSTSUBSCRIPT p end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT p end_POSTSUBSCRIPT when σc=0subscript𝜎c0\sigma_{\textrm{c}}=0italic_σ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = 0, although, in practice, σcsubscript𝜎c\sigma_{\textrm{c}}italic_σ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT and σpsubscript𝜎p\sigma_{\textrm{p}}italic_σ start_POSTSUBSCRIPT p end_POSTSUBSCRIPT never attain zero values since they are bounded below with limiting values.

We refer to the hybridized value (σmsubscript𝜎m\sigma_{\textrm{m}}italic_σ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT) as the modelled background-error standard deviation. For the parameterized standard deviations, we set wp=1subscript𝑤p1w_{\textrm{p}}=1italic_w start_POSTSUBSCRIPT p end_POSTSUBSCRIPT = 1 for all variables. For the climatological standard deviations, we set wc=1subscript𝑤c1w_{\textrm{c}}=1italic_w start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = 1 for T𝑇Titalic_T and Susubscript𝑆uS_{\rm u}italic_S start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT, and wc=0.5subscript𝑤c0.5w_{\textrm{c}}=0.5italic_w start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = 0.5 for ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT. The climatological standard deviations are down-weighted for ηusubscript𝜂u\eta_{\rm u}italic_η start_POSTSUBSCRIPT roman_u end_POSTSUBSCRIPT as we found that applying them at the full magnitude makes the assimilation system draw too strongly to the altimeter observations, which results in degraded performance. Subsequently, the modelled standard deviations, σmsubscript𝜎m\sigma_{\textrm{m}}italic_σ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT, are combined, also using the smooth maximum function, with “errors of the day”, σesubscript𝜎e\sigma_{\textrm{e}}italic_σ start_POSTSUBSCRIPT e end_POSTSUBSCRIPT, estimated from the forecast ensemble:

σh=1hlog(ehwmσm+ehweσe1)subscript𝜎h1superscript𝑒subscript𝑤msubscript𝜎msuperscript𝑒subscript𝑤esubscript𝜎e1\sigma_{\textrm{h}}=\frac{1}{h}\log\left(e^{hw_{{\textrm{m}}}\sigma_{\textrm{m% }}}+e^{hw_{\textrm{e}}\sigma_{\textrm{e}}}-1\right)italic_σ start_POSTSUBSCRIPT h end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_h end_ARG roman_log ( italic_e start_POSTSUPERSCRIPT italic_h italic_w start_POSTSUBSCRIPT m end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT italic_h italic_w start_POSTSUBSCRIPT e end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT e end_POSTSUBSCRIPT end_POSTSUPERSCRIPT - 1 ) (14)

where wmsubscript𝑤mw_{\textrm{m}}italic_w start_POSTSUBSCRIPT m end_POSTSUBSCRIPT and wesubscript𝑤ew_{\textrm{e}}italic_w start_POSTSUBSCRIPT e end_POSTSUBSCRIPT are weighting factors for the modelled and ensemble- derived standard deviations, respectively. We refer to σhsubscript𝜎h\sigma_{\textrm{h}}italic_σ start_POSTSUBSCRIPT h end_POSTSUBSCRIPT as the hybrid (modelled–flow-dependent) standard deviations. Similar to the hybridization of the climatological component in Equation (13), we set wm=we=1subscript𝑤msubscript𝑤e1w_{\textrm{m}}=w_{\textrm{e}}=1italic_w start_POSTSUBSCRIPT m end_POSTSUBSCRIPT = italic_w start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = 1 for T𝑇Titalic_T and Susubscript𝑆uS_{\textrm{u}}italic_S start_POSTSUBSCRIPT u end_POSTSUBSCRIPT, wm=1subscript𝑤m1w_{\textrm{m}}=1italic_w start_POSTSUBSCRIPT m end_POSTSUBSCRIPT = 1 and we=0.5subscript𝑤e0.5w_{\textrm{e}}=0.5italic_w start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = 0.5 for ηusubscript𝜂u\eta_{\textrm{u}}italic_η start_POSTSUBSCRIPT u end_POSTSUBSCRIPT, and used the same values of hhitalic_h.

Refer to caption Refer to caption
Refer to caption Refer to caption
Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 1: Rows from the top: climatological (estimated from ensemble perturbations for months December-January-February (DJF) in the period 2010–2015), “errors of the day”, parameterized and hybrid standard deviations (SD) for temperature at 97 m depth and the unbalanced component of sea surface height (SSHu) (right column) for the cycle starting on 04/01/2017.

As described above, the standard deviations specified in the new system are obtained by combining, using Equation 14, the legacy ORAS5 parameterized standard deviations with climatological standard deviations and “errors of the day” diagnosed from the ensemble. This is illustrated Figure 1 for a particular cycle, valid on January 4thsuperscript4th4^{\textrm{th}}4 start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT, 2017201720172017, for temperature, in level 24 of the ocean model, and for unbalanced SSH. The climatological background-error standard deviations are illustrated in the top row for the boreal winter (December-January-February). Not surprisingly, the errors are largest in eddy-active regions and rather small otherwise. This is well visible for the unbalanced SSH standard deviations, which appear to be close to zero apart from the WBC and ACC regions. Seasonal variability, not shown in the figure, is generally small. It can be best observed for the Gulf Stream and Kuroshio Current regions, where the errors are larger in the boreal winter and relatively smaller in the boreal summer. The standard deviations corresponding to the “errors of the day”, illustrated in the second row, show sharp flow-dependent features in eddy-active regions. Kelvin wave-like structures are visible in the tropical Pacific for temperature. On the other hand, elsewhere the standard deviations are suppressed compared to the parameterized values. This is particularly visible for unbalanced SSH. It is interesting to contrast the “errors of the day” with the climatological values. The structures are, as expected, much smoother. The location of the largest standard deviations is unsurprisingly very similar and confined to eddy-active regions The parameterized standard deviations, illustrated in the thrid row of the figure, show a limited degree of flow dependency in the extra-tropics in the case of temperature and are constant for unbalanced SSH. It should be noted that in the case of unbalanced SSH, the standard deviations are smoothly damped to zero in the tropics. Finally, the last row shows the combined hybrid standard deviations. Owing to the properties of the smooth maximum function, the structure resembles that of sharpened climatological standard deviations with the background magnitude outside of eddy-active regions corresponding to that of the parameterized values.

Figure 2 shows the vertical structure of the ensemble and parameterized background-error standard deviations at the Equator (top row) and at the constant latitude of 40superscript4040^{\circ}40 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTN (bottom row), valid on January 4thsuperscript4th4^{\textrm{th}}4 start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT, 2017. Compared to the parameterized values, the ensemble-derived parameters show similar vertical structure at the Equator, but the parameterized values are substantially larger in the mixed layer and at the level of the thermocline. At the constant latitude of 40superscript4040^{\circ}40 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTN, the ensemble-derived parameters show richer vertical and horizontal structure, and are larger in the mixed layer in the Gulf Stream and Kuroshio Current regions and smaller elsewhere apart from the deep ocean where both are close to zero.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 2: Vertical section of the background-error standard deviations (SD) for temperature estimated from the ensemble (left panel) and from the ORAS5 parameterization (right panel) at the Equator (top row) and at a constant latitude of 40superscript4040^{\circ}40 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTN (bottom row), valid on 04/01/2017.

Examining the new hybrid background-error standard deviations, it is clear that the background forecast is given less confidence in the eddy-active regions in the new system and as a consequence is allowed to draw more to observations there. Such behaviour is desirable in order to improve the accuracy of the initial state where the uncertainty in the background state is highest.

2.3.2 Hybrid diffusion tensor

As described in Section 2.1, the implicit diffusion operator used to model the correlation matrix requires normalization. Computing accurate normalization factors for a 3D implicit diffusion operator is costly, which precludes us from using a fully flow-dependent 3D correlation model. When the horizontal and vertical correlations are modelled separately as in Equation (3), Weaver et al. (2021) showed that the normalization matrix can be well approximated by a product of a normalization matrix for the horizontal diffusion component and a normalization matrix for the vertical diffusion component, where each can be estimated separately using randomisation or exactly in the case of the latter. This approximation allows us to decouple the specification of the horizontal and vertical diffusion tensor components. While the cost of computing the normalization factors of the horizontal component is high and can be considered not affordable in a cycling system, the cost of computing the vertical normalization factors is low. For this reason, the new system uses a climatological horizontal diffusion tensor and a fully flow-dependent vertical diffusion tensor.

Figure 3 shows a comparison of the zonal and meridional components of the horizontal correlation length-scales for temperature at 1111 m depth between ORAS5 in the left column and the new system in the right column. In ORAS5, the horizontal correlation length-scales, which are proportional to the square-root of the diagonal elements of the horizontal diffusion tensor, were specified for temperature using a parameterization based on the Rossby radius. The correlation length-scales are largest in the tropics and decrease with increasing latitude. Zonal correlation length-scales are larger than meridional correlation length-scales, reflecting well known anisotropy related to equatorial dynamics. The anisotropy is also clearly visible in the climatological length-scales. Similarly, the length-scales are largest in the tropics, albeit shorter than in ORAS5, and smaller in the extra-tropics, in particular in the WBC and ACC regions. In the tropics, the climatological length-scales show clear minima associated with the location of Tropical Instability Waves. It is also interesting to observe that larger correlation length-scales are diagnosed in the Southern Hemisphere in the boreal winter, which is not the case in boreal summer (not shown).

Figure 4 shows the comparison of the zonal and meridional components of the horizontal correlation length-scales for unbalanced SSH. As can be seen in the left column of the figure, the zonal and meridional length-scales are equal and constant everywhere, except in the proximity of land boundaries where they are reduced. The climatological length-scales are significantly shorter and also reduced artificially in the proximity of land boundaries. Similarly, as for temperature, the zonal correlation length-scales are larger than the meridional correlation length-scales, reflecting the anisotropic nature of the dynamics in this region.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 3: Temperature meridional (L11, top row) and zonal (L22, bottom row) horizontal background-error correlation length-scales at 1 m depth. Left: ORAS5 parameterized length-scales. Right: climatological length-scales estimated from ensemble perturbations for months December-January-February (DJF) in the period 2010–2015.
Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 4: Unbalanced component of SSH (SSHu) meridional (L11, top row) and zonal (L22, bottom row) horizontal background-error correlation length-scales. Left: ORAS5 parameterized length-scales. Right: climatological length-scales estimated from ensemble perturbations from months December-January-February (DJF) in the period 2010–2015.

The vertical diffusion tensor coefficients are parameterized in ORAS5 such that the vertical correlation length-scales are proportional to the vertical grid-scale factors. The proportionality constant was set to be equal to one. The vertical correlation length-scales estimated from the ensemble are shown in Figure 5 for the assimilation cycle starting on January 4thsuperscript4th4^{\textrm{th}}4 start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT, 2017. The top row panel shows the vertical correlation length-scales at 97 m depth. The vertical length-scales near the surface are much larger in the Northern Hemisphere than in the Southern Hemisphere. This is consistent with the expectation that the errors are strongly correlated in the mixed layer, which is deeper in the boreal winter in the Northern Hemisphere. The bottom row panels show a vertical cross section of the vertical correlation length-scales at the Equator (left panel) and at the 40superscript4040^{\circ}40 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTN latitude (right panel). At 40superscript4040^{\circ}40 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTN latitude, the highly stratified thermocline region, where the vertical length-scales are markedly shorter, can be clearly distinguished from both the mixed layer and the deep ocean, where the length-scales are much larger. The parameterized formulation of the vertical diffusion tensor used in ORAS5, by construction, neither distinguishes the thermocline region, nor well represents the length-scales in the mixed layer and the deep ocean. In both cases, the vertical length-scales are much smaller than those estimated from the ensemble. In fact, the short vertical correlation length-scales in the mixed layer in the parameterized error covariances make the assimilation of SST ineffective, as will be illustrated in the next section.

Refer to caption
Refer to caption Refer to caption
Figure 5: Top row: vertical (L33) background-error correlation length-scales estimated from the ensemble for temperature at 97 m depth. Bottom row: vertical section at the Equator (left panel) and at the constant latitude of 40superscript4040^{\circ}40 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTN (right panel) of the vertical (L33) background-error correlation length-scales estimated from the ensemble. Plots are valid on 04/01/2017.

3 Results

In this section, we demonstrate the impact of the new, hybrid formulation of 𝐁𝐁\mathbf{B}bold_B. We carry out a series of experiments in the observation-rich period spanning 2017–2021. The evaluation period is outside the period which we used to compute the climatology of the parameters of 𝐁𝐁\mathbf{B}bold_B. Before commenting on the results, we start by describing our baseline configuration in Section 3.1. Rather than presenting the overall impact of the new, hybrid 𝐁𝐁\mathbf{B}bold_B formulation, we first show the impact of the hybrid variances and the hybrid tensor in isolation in Section 3.2 and Section 3.3, respectively. Finally, the impact of the complete hybrid 𝐁𝐁\mathbf{B}bold_B formulation is presented in Section 3.4. Our evaluation metric focuses on analysing the impact on Root Mean Squared (RMS) errors of the background forecasts relative to the observations (before assimilation). We consider the new system to be an improvement over the baseline configuration whenever the background errors are reduced relative to those from the baseline configuration. We do not attempt in this article to analyse the impact on climate trends. This topic will be addressed in a separate article.

3.1 Baseline configuration

Our baseline experiment for evaluating the impact of the new, hybrid 𝐁𝐁\mathbf{B}bold_B formulation employs the strictly parameterized formulation of 𝐁𝐁\mathbf{B}bold_B that was used for ORAS5. The assimilation window is 5555 days like in ORAS5. The assimilated observations include information on temperature and salinity from in situ measurements (Argo floats, moored buoys, ships, marine mammals, and gliders), sea level anomalies (SLA) from altimeters, and sea ice concentration from L3 satellite-based products. Unlike in ORAS5, the L4 OSTIA SST data are assimilated in NEMOVAR. All other aspects of the system, including the model bias correction scheme and the mean dynamic topography used for referencing SLA observations are those developed for ORAS6. Our goal is to evaluate the impact of the hybrid 𝐁𝐁\mathbf{B}bold_B formulation with respect to its parameterized predecessor.

3.2 Impact of the hybrid variances

We focus first on evaluating the impact of replacing the parameterized variances with the newly developed hybrid variances that we described in Section 2.3.1. The diffusion tensor is the same parameterized tensor used in the baseline experiment. Figure 6 shows spatial maps of the relative change in RMS background errors for temperature and salinity averaged over the two depth ranges of 0–200 m and 200–1000 m. For temperature, the results are largely positive in the extra-tropics, in particular in regions where the climatological parameters are able to describe more faithfully the ocean variability, specifically in the WBC and ACC regions. Some degradation is visible in the tropics, both in the Atlantic Ocean and Pacific Ocean. The hybrid variances, which by construction are larger than the parameterized values, allow the system to draw closely to observations in these regions. The results for salinity are more mixed. The RMS errors are reduced in the extra-tropics, but are also increased in the tropical Pacific Ocean and tropical Atlantic Ocean. There is a sign of degradation in the Antarctic region too. Figure 7 shows the relative change in the standard deviation of background errors for SSH. While the errors are significantly reduced in the WBC and ACC regions, there is also large degradation in the tropics and south of the ACC. As can be seen, using the hybrid variances in isolation brings a mixed bag of results. While the overall performance is positive, especially in the extra-tropics, noticeable degradation is also present. It should be noted that the hybrid variances were computed together with the hybrid tensor and are thus representative of the error structures induced by the latter. Still, we find it of interest to assess the impact of hybrid variances in isolation in order to better understand where the improvements associated with the hybrid 𝐁𝐁\mathbf{B}bold_B come from. In particular, such results may prove useful when trying to explain the origins of degradation, should there be any, of the full hybrid 𝐁𝐁\mathbf{B}bold_B.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 6: Top row: impact of using hybrid variances in 𝐁𝐁\mathbf{B}bold_B on the temperature RMS background errors in the top 200 m (left panel) and in the depth-range 200–1000 m (right panel); bottom row: the same but for salinity. The errors are plotted relative to those of the baseline experiment, which employed a parameterized 𝐁𝐁\mathbf{B}bold_B.
Refer to caption
Figure 7: Impact of using hybrid variances in 𝐁𝐁\mathbf{B}bold_B on the standard deviation of SSH background errors. The errors are plotted relative to those of the baseline experiment, which employed a parameterized 𝐁𝐁\mathbf{B}bold_B.

3.3 Impact of the hybrid tensor

As described in Section 2.3.2, the hybrid correlation matrix is built from a horizontal diffusion operator equipped with a climatological diffusion tensor and a vertical diffusion operator equipped with a fully flow-dependent vertical diffusion tensor. The vertical correlation length-scales diagnosed from the forecast ensemble are typically much larger in the mixed layer and in the deep ocean compared to those from the parameterized formulation. The larger values in the mixed layer should in principle allow the SST observations to have larger impact on the analysis by projecting them further into the mixed layer. Figure 8 shows spatial maps of the relative change in RMS background errors for temperature and salinity for the two depth ranges of 0–200m and 200–1000m. For temperature in the extra-tropics, we can see a similar pattern to that obtained with the hybrid variances, where the errors are significantly reduced in the eddy-active regions. Interestingly, the background errors are now also reduced in the tropics, in particular in the upper ocean. Unlike when using hybrid variances, the background errors are larger in Baffin Bay at both plotted depth ranges. Salinity forecast errors are reduced in the tropics, with the exception of the Banda Sea, and the high-latitude extra-tropics. There is significant degradation in Baffin Bay and in the Arctic Ocean. Some degradation is also visible off the coasts of Antarctica. Figure 9 shows the relative change in the standard deviation of background errors for SSH. The background errors are significantly reduced everywhere except in Baffin Bay and the Arctic Ocean. The degradation in all three variables at high latitudes may be associated with the relatively small horizontal climatological correlation length-scales compared to the parameterized length-scales, which is evident in Figure 3 and Figure 4, for temperature and SSH, respectively.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 8: Top row: impact of using the hybrid diffusion tensor in 𝐁𝐁\mathbf{B}bold_B on the temperature RMS background errors in the top 200 m (left panel) and in the depth-range 200–1000m (right panel); bottom row: the same but for salinity. The errors are plotted relative to those of the baseline experiment, which employed a parameterized 𝐁𝐁\mathbf{B}bold_B.
Refer to caption
Figure 9: Impact of using the hybrid diffusion tensor in 𝐁𝐁\mathbf{B}bold_B on the standard deviation of SSH background errors. The errors are plotted relative to those of the baseline experiment, which employed a parameterized 𝐁𝐁\mathbf{B}bold_B.

3.4 Impact of the hybrid variances and hybrid tensor combined

Refer to caption
Figure 10: Globally averaged normalized RMS errors for the analysis (left panel) and background (right panel) computed with respect to the assimilated observations for temperature as a function of ocean depth. The baseline experiment with the parameterized formulation of 𝐁𝐁\mathbf{B}bold_B marks the 100% line.
Refer to caption
Figure 11: Globally averaged normalized RMS errors for the analysis (left panel) and background (right panel) computed with respect to the assimilated observations for salinity as a function of ocean depth. The baseline experiment with the parameterized 𝐁𝐁\mathbf{B}bold_B marks the 100% line.
Refer to caption
Figure 12: Globally averaged normalized RMS errors for the analysis (left panel) and background (right panel) computed with respect to the assimilated observations for SST and SLA. The baseline experiment with the parameterized formulation of 𝐁𝐁\mathbf{B}bold_B marks the 100% line.

We demonstrated in the previous subsections that the impact of the hybrid variances and the hybrid tensor on reducing background errors is generally positive when each of them is tested independently from the other. There were some caveats, however. In both cases, the impact is largely positive in the WBC and ACC regions, which are important regions from the point of view of medium-range coupled ocean-atmosphere forecasting. We now check the performance of using the hybrid variances together with the hybrid tensor.

In order to provide a more quantitative assessment of the impact, we show in Figure 10 the globally averaged vertical profiles of the normalized RMS errors of the analysis and background computed against the assimilated observations for temperature for the three considered configurations: with hybrid variances, hybrid diffusion tensor and both together. The RMS errors are normalized with the RMS errors of the baseline experiment employing the parameterized 𝐁𝐁\mathbf{B}bold_B. As could be anticipated, applying hybrid variances while keeping the parameterized tensor results in an analysis that fits much more closely the observations. This is consistent with the fact that the climatological variances are larger in the eddy-active regions around the globe. The improvement of the background fit to observations, on the other hand, is relatively modest, especially in the upper ocean. This may be due either to over-fitting the observations or mis-specifying the correlation length-scales. When applying the hybrid tensor together with parameterized variances, on the other hand, the background fit to the observations is improved to a similar extent as the analysis fit to the observations. It may be expected that both drawing more to observations in the analysis and improving the representation of the structure of the correlation matrix should be beneficial. This is indeed the case as the configuration with both hybrid variances and hybrid tensor performs the best. The globally averaged temperature RMS errors are reduced by roughly 10%percent1010\%10 % throughout the whole water column with respect to the configuration employing the ORAS5 𝐁𝐁\mathbf{B}bold_B.

Figure 11 shows the corresponding globally averaged vertical profiles of normalized RMS errors for salinity. This time, using the hybrid tensor alone results in a degraded analysis fit to observations in the upper 1000 m, while the background fit is only marginally improved. Employing hybrid variances results in the analysis drawing closer to observations, but with little impact on the background fit to observations. Only when the hybrid variances are combined with the hybrid tensor, can we see a significant improvement to the system’s performance. The background fit to salinity observations is improved by over 5%percent55\%5 % in the upper ocean and up to 10%percent1010\%10 % around a depth of 600 m. Figure 12 compares the globally averaged analysis and background fit to SST and SLA observations for the three configurations. Once again, while both the hybrid variance and hybrid tensor configurations show reduced analysis and background errors, it is when combining both hybrid elements that we see the best performance. Both the background fit to SST and SLA observations is improved by nearly 20%percent2020\%20 %.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 13: Top row: impact of using the hybrid 𝐁𝐁\mathbf{B}bold_B on the temperature RMS background errors in the top 200 m (left panel) and in the depth-range 200–1000 m (right panel); bottom row: the same but for salinity.
Refer to caption
Figure 14: Impact of using the hybrid tensor in 𝐁𝐁\mathbf{B}bold_B on the standard deviation of the background errors for SSH.

Figure 13 shows spatial maps of the relative RMS background errors for temperature and salinity for the two depth ranges of 0–200m and 200–1000m. The temperature RMS errors are significantly reduced in the WBC and ACC regions. The degradation in Baffin Bay, also observed when using the hybrid tensor in isolation, is also visible. The salinity background errors are significantly reduced in the upper ocean apart from high-latitude regions in the Northern and Southern Hemispheres, which, as remarked earlier, may be associated with shorter horizontal correlation length-scales in the hybrid tensor. Figure 14 shows the relative change in the standard deviation of background errors for SSH. Apart from impressive background-error reduction in eddy-active regions, Baffin Bay stands out as an area where the background errors are increased.

The location of large reductions in the background errors correlates well with areas where the hybrid variances are larger compared to their parameterized counterparts. As could be appreciated from figures showing vertical profiles of RMS analysis and background errors, the hybrid variances only explain part of the improvements. It is the hybrid tensor, with better representation of correlation spatial structure, that brings a significant impact. While the horizontal correlation length-scales are better aligned with the boundaries of ocean currents, the flow-dependent specification of vertical correlation length-scales plays an even more significant role, in particular in improving the vertical projection of information coming from SST observations. The increments obtained when using the parameterized formulation of 𝐁𝐁\mathbf{B}bold_B are relatively shallow and therefore incapable of correcting background errors throughout the mixed layer. The flow-dependent specification of vertical correlation length-scales, on the other hand, results in increments extending to the bottom of the mixed layer and thus provides a more effective correction to the upper ocean.

The inability of the parameterized 𝐁𝐁\mathbf{B}bold_B to project the SST increments into the mixed layer causes large model biases to develop in a cycled data assimilation system. This problem is particularly acute in the Gulf Stream area. Since IFS cycle 45r1, ECMWF’s medium-range forecasts are fully coupled to OCEAN5 in the tropics, but are only partially coupled in the extra-tropics (Browne et al., 2019). This partial coupling strategy is used because of a known deficiency of ORAS5: it is not able to accurately represent the position of western boundary currents. Figure 15 shows the SST bias computed for the period 2017–2021 when using the parameterized 𝐁𝐁\mathbf{B}bold_B (left panel) and when using the new hybrid 𝐁𝐁\mathbf{B}bold_B. The bias is largely reduced, thus mitigating the need for a partial coupling strategy in ORAS6.

Refer to caption Refer to caption
Figure 15: SST bias (in K) obtained when using the parameterized 𝐁𝐁\mathbf{B}bold_B (left panel) and the new hybrid 𝐁𝐁\mathbf{B}bold_B (right panel). SST is verified against the European Space Agency CCI2 SST dataset in the period 2017–2021.

4 Conclusions

In this article, we have described an Ensemble of Data Assimilations framework that has been developed for ECMWF’s next generation ocean reanalysis system, ORAS6. The forecast ensemble is used in two ways: first, to construct seasonal climatologies of parameters of the background-error covariance matrix (𝐁𝐁{\bf B}bold_B), specifically the variances and the horizontal and vertical length-scales of the correlation model; and second, to provide flow-dependent estimates, on each assimilation cycle, of the background-error variances and vertical correlation length-scales. The climatological and flow-dependent estimates are blended in a hybrid formulation of 𝐁𝐁\mathbf{B}bold_B. The impact of this new, hybrid 𝐁𝐁\mathbf{B}bold_B has been assessed with respect to the formulation of 𝐁𝐁\mathbf{B}bold_B used in OCEAN5 and ORAS5, ECMWF’s current ocean analysis and reanalysis systems. In the current system, the variances and correlation length-scales are parameterized using empirical and analytical relationships, without making use of an ensemble.

The performance of the new formulation of 𝐁𝐁\mathbf{B}bold_B has been assessed in the Argo-rich period. The climatological estimates of the covariance parameters have also been computed in the Argo-rich period, over a 6-year interval preceding the assessment period. The new formulation of 𝐁𝐁\mathbf{B}bold_B has not been assessed, however, in periods prior to the introduction of Argo. It is worth remarking that the climatological estimates of the covariance parameters computed here are not appropriate for the pre-Argo period when background errors can be expected to be much larger. Applying the hybrid 𝐁𝐁\mathbf{B}bold_B in that period, or in any period where there is a substantial change in the observation network, should require carefully revising the climatology.

The results from this study are overwhelmingly positive with the new, hybrid formulation of 𝐁𝐁\mathbf{B}bold_B, with background errors being significantly reduced for temperature, salinity and sea surface height. Improvements are largest in extra-tropical eddy-active regions. Relatively small degradation is present at high latitudes, in particular in Baffin Bay and in the Arctic Ocean. An important feature of the new 𝐁𝐁\mathbf{B}bold_B is that it allows increments from SST observations to be projected effectively into the mixed layer, which is not possible with the parameterized 𝐁𝐁\mathbf{B}bold_B used in ORAS5. As a result, in ORAS6, SST observations can now be assimilated directly using the 3D-Var FGAT scheme, and hence in synergy with the other assimilated observations, instead of indirectly via an ad hoc nudging scheme as done in ORAS5.

There is considerable scope for improving the EDA, both in the ensemble-generation strategy and in the way the ensembles are used in the hybrid 𝐁𝐁\mathbf{B}bold_B. Improving the quality and reliability of the forecast ensemble can potentially be achieved by introducing stochastic perturbations in the ocean model to account for uncertainty arising from random model errors. The covariance model can also be enhanced to make more effective use of the ensemble information. For example, the diffusion-based correlation model could be generalized to account for a non-diagonal diffusion tensor in order to exploit the anisotropic information in the ensemble more appropriately than done here. Building scale dependence into the covariance model, so that observations can correct different spatial scales more effectively, is another promising avenue to explore. Recent work to develop a scale-dependent, ensemble-based 𝐁𝐁\mathbf{B}bold_B suggests that this approach can be substantially beneficial to the analysis.

Acknowledgements

This work has benefited from funding from the Copernicus Climate Change Service (C3S), one of six services of the European Union’s Copernicus Programme, which is implemented by ECMWF on behalf of the European Commission. In particular, this work is a contribution to contract C3S_321b_INRIA: “Technical preparations for C3S Seasonal Initialization and Global Reanalysis: Enabling an Ensemble of Data Assimilations for the Ocean”. We would like to thank Massimo Bonavita and Tony McNally for their comments and suggestions to improve the manuscript.

References

  • Balmaseda et al. (2013) Balmaseda, M. A., Mogensen, K. and Weaver, A. T. (2013). Evaluation of the ECMWF ocean reanalysis system ORAS4. Quarterly Journal of the Royal Meteorological Society, 139 (674), 1132–1161. doi:https://doi.org/10.1002/qj.2063.
    URL https://rmets.onlinelibrary.wiley.com/doi/abs/10.1002/qj.2063
  • Berre et al. (2015) Berre, L., Varella, H. and Desroziers, G. (2015). Modelling of flow-dependent ensemble-based background-error correlations using a wavelet formulation in 4D-Var at Météo-France. Quarterly Journal of the Royal Meteorological Society, 141 (692), 2803–2812. doi:https://doi.org/10.1002/qj.2565.
    URL https://rmets.onlinelibrary.wiley.com/doi/abs/10.1002/qj.2565
  • Bloom et al. (1996) Bloom, S. C., Takacs, L. L., da Silva, A. M. and Ledvina, D. (1996). Data assimilation using incremental analysis updates. Monthly Weather Review, 124 (6), 1256 – 1271. doi:10.1175/1520-0493(1996)124<1256:DAUIAU>2.0.CO;2.
    URL https://journals.ametsoc.org/view/journals/mwre/124/6/1520-0493_1996_124_1256_dauiau_2_0_co_2.xml
  • Bonavita et al. (2016) Bonavita, M., Hólm, E., Isaksen, L. and Fisher, M. (2016). The evolution of the ECMWF hybrid data assimilation system. Quarterly Journal of the Royal Meteorological Society, 142 (694), 287–303. doi:https://doi.org/10.1002/qj.2652.
    URL https://rmets.onlinelibrary.wiley.com/doi/abs/10.1002/qj.2652
  • Bonavita et al. (2012) Bonavita, M., Isaksen, L. and Hólm, E. (2012). On the use of EDA background error variances in the ECMWF 4D-Var. Quarterly Journal of the Royal Meteorological Society, 138 (667), 1540–1559. doi:https://doi.org/10.1002/qj.1899.
    URL https://rmets.onlinelibrary.wiley.com/doi/abs/10.1002/qj.1899
  • Browne et al. (2019) Browne, P. A., de Rosnay, P., Zuo, H., Bennett, A. and Dawson, A. (2019). Weakly coupled ocean–atmosphere data assimilation in the ECMWF NWP system. Remote Sensing, 11 (3). doi:10.3390/rs11030234.
    URL https://www.mdpi.com/2072-4292/11/3/234
  • Buehner (2005) Buehner, M. (2005). Ensemble-derived stationary and flow-dependent background-error covariances: Evaluation in a quasi-operational NWP setting. Quarterly Journal of the Royal Meteorological Society, 131 (607), 1013–1043. doi:10.1256/qj.04.15.
    URL https://doi.org/10.1256/qj.04.15
  • Clayton et al. (2013) Clayton, A. M., Lorenc, A. C. and Barker, D. M. (2013). Operational implementation of a hybrid ensemble/4D-Var global data assimilation system at the Met Office. Quarterly Journal of the Royal Meteorological Society, 139 (675), 1445–1461. doi:https://doi.org/10.1002/qj.2054.
    URL https://rmets.onlinelibrary.wiley.com/doi/abs/10.1002/qj.2054
  • Daget et al. (2009) Daget, N., Weaver, A. T. and Balmaseda, M. A. (2009). Ensemble estimation of background-error variances in a three-dimensional variational data assimilation system for the global ocean. Quarterly Journal of the Royal Meteorological Society, 135 (641), 1071–1094. doi:https://doi.org/10.1002/qj.412.
    URL https://rmets.onlinelibrary.wiley.com/doi/abs/10.1002/qj.412
  • Daley (1991) Daley, R. (1991). Atmospheric Data Analysis. Cambridge Atmospheric and Space Sciences Series. Cambridge, UK: Cambridge University Press.
  • de Rosnay et al. (2022) de Rosnay, P., Browne, P., de Boisséson, E., Fairbairn, D., Hirahara, Y., Ochi, K., Schepers, D., Weston, P., Zuo, H., Alonso-Balmaseda, M., Balsamo, G., Bonavita, M., Borman, N., Brown, A., Chrust, M., Dahoui, M., Chiara, G., English, S., Geer, A., Healy, S., Hersbach, H., Laloyaux, P., Magnusson, L., Massart, S., McNally, A., Pappenberger, F. and Rabier, F. (2022). Coupled data assimilation at ECMWF: current status, challenges and future developments. Quarterly Journal of the Royal Meteorological Society, 148 (747), 2672–2702. doi:https://doi.org/10.1002/qj.4330.
    URL https://rmets.onlinelibrary.wiley.com/doi/abs/10.1002/qj.4330
  • Goux et al. (2024) Goux, O., Gürol, S., Weaver, A. T., Diouane, Y. and Guillet, O. (2024). Impact of correlated observation errors on the conditioning of variational data assimilation problems. Numerical Linear Algebra with Applications, 31 (1), e2529. doi:https://doi.org/10.1002/nla.2529.
    URL https://onlinelibrary.wiley.com/doi/abs/10.1002/nla.2529
  • Guttorp and Gneiting (2006) Guttorp, P. and Gneiting, T. (2006). Studies in the history of probability and statistics XLIX on the Matérn correlation family. Biometrika, 93, 989–95.
  • Hersbach et al. (2020) Hersbach, H., Bell, B., Berrisford, P., Hirahara, S., Horányi, A., Muñoz-Sabater, J., Nicolas, J., Peubey, C., Radu, R., Schepers, D., Simmons, A., Soci, C., Abdalla, S., Abellan, X., Balsamo, G., Bechtold, P., Biavati, G., Bidlot, J., Bonavita, M., Chiara, G., Dahlgren, P., Dee, D., Diamantakis, M., Dragani, R., Flemming, J., Forbes, R., Fuentes, M., Geer, A., Haimberger, L., Healy, S., Hogan, R. J., Hólm, E., Janisková, M., Keeley, S., Laloyaux, P., Lopez, P., Lupu, C., Radnoti, G., Rosnay, P., Rozum, I., Vamborg, F., Villaume, S. and Thépaut, J.-N. (2020). The era5 global reanalysis. Quarterly Journal of the Royal Meteorological Society, 146 (730), 1999–2049. doi:10.1002/qj.3803.
  • Kleist and Ide (2015) Kleist, D. T. and Ide, K. (2015). An OSSE-Based Evaluation of Hybrid Variational–Ensemble Data Assimilation for the NCEP GFS. Part I: System Description and 3D-Hybrid Results. Monthly Weather Review, 143 (2), 433 – 451. doi:10.1175/MWR-D-13-00351.1.
    URL https://journals.ametsoc.org/view/journals/mwre/143/2/mwr-d-13-00351.1.xml
  • Lea et al. (2022) Lea, D. J., While, J., Martin, M. J., Weaver, A., Storto, A. and Chrust, M. (2022). A new global ocean ensemble system at the Met Office: assessing the impact of hybrid data assimilation and inflation settings. Quarterly Journal of the Royal Meteorological Society, 148 (745), 1996–2030. doi:https://doi.org/10.1002/qj.4292.
    URL https://rmets.onlinelibrary.wiley.com/doi/abs/10.1002/qj.4292
  • Lorenc (2003) Lorenc, A. C. (2003). The potential of the ensemble Kalman filter for NWP—a comparison with 4D-Var. Quarterly Journal of the Royal Meteorological Society, 129 (595), 3183–3203. doi:10.1256/qj.02.132.
    URL https://doi.org/10.1256/qj.04.15
  • Madec et al. (2023) Madec, G., Bell, M., Blaker, A., Bricaud, C., Bruciaferri, D., Castrillo, M., Calvert, D., Chanut, J., Clementi, E., Coward, A., Epicoco, I., Éthé, C., Ganderton, J., Harle, J., Hutchinson, K., Iovino, D., Lea, D., Lovato, T., Martin, M., Martin, N., Mele, F., Martins, D., Masson, S., Mathiot, P., Mele, F., Mocavero, S., Müller, S., Nurser, A. G., Paronuzzi, S., Peltier, M., Person, R., Rousset, C., Rynders, S., Samson, G., Téchené, S., Vancoppenolle, M. and Wilson, C. (2023). NEMO Ocean Engine Reference Manual.
    URL https://doi.org/10.5281/zenodo.8167700
  • Michel et al. (2016) Michel, Y., Ménétrier, B. and Montmerle, T. (2016). Objective filtering of the local correlation tensor. Quarterly Journal of the Royal Meteorological Society, 142.
    URL https://api.semanticscholar.org/CorpusID:124391666
  • Mogensen et al. (2012) Mogensen, K., Balmaseda, M. A. and Weaver, A. T. (2012). The NEMOVAR ocean data assimilation system as implemented in the ECMWF ocean analysis for System 4.
    URL https://www.ecmwf.int/node/11174
  • Ménétrier et al. (2015) Ménétrier, B., Montmerle, T., Michel, Y. and Berre, L. (2015). Linear filtering of sample covariances for ensemble-based data assimilation. Part I: optimality criteria and application to variance filtering and covariance localization. Monthly Weather Review, 143 (5), 1622 – 1643. doi:10.1175/MWR-D-14-00157.1.
    URL https://journals.ametsoc.org/view/journals/mwre/143/5/mwr-d-14-00157.1.xml
  • Penny et al. (2015) Penny, S. G., Behringer, D. W., Carton, J. A. and Kalnay, E. (2015). A hybrid global ocean data assimilation system at NCEP. Monthly Weather Review, 143 (11), 4660–4677. doi:10.1175/MWR-D-14-00376.1.
  • Ricci et al. (2005) Ricci, S., Weaver, A. T., Vialard, J. and Rogel, P. (2005). Incorporating state-dependent temperature–salinity constraints in the background error covariance of variational ocean data assimilation. Monthly Weather Review, 133 (1), 317 – 338. doi:10.1175/MWR2872.1.
    URL https://journals.ametsoc.org/view/journals/mwre/133/1/mwr2872.1.xml
  • Storto et al. (2018) Storto, A., Oddo, P., Cipollone, A., Mirouze, I. and Lemieux-Dudon, B. (2018). Extending an oceanographic variational scheme to allow for affordable hybrid and four-dimensional data assimilation. Ocean Modelling, 128, 67–86. doi:https://doi.org/10.1016/j.ocemod.2018.06.005.
    URL https://www.sciencedirect.com/science/article/pii/S1463500318302282
  • Weaver et al. (2021) Weaver, A. T., Chrust, M., Ménétrier, B. and Piacentini, A. (2021). An evaluation of methods for normalizing diffusion-based covariance operators in variational data assimilation. Quarterly Journal of the Royal Meteorological Society, 147 (734), 289–320. doi:https://doi.org/10.1002/qj.3918.
    URL https://rmets.onlinelibrary.wiley.com/doi/abs/10.1002/qj.3918
  • Weaver et al. (2005) Weaver, A. T., Deltel, C., Machu, E., Ricci, S. and Daget, N. (2005). A multivariate balance operator for variational ocean data assimilation. Quarterly Journal of the Royal Meteorological Society, 131 (613), 3605–3625. doi:https://doi.org/10.1256/qj.05.119.
    URL https://rmets.onlinelibrary.wiley.com/doi/abs/10.1256/qj.05.119
  • Weaver and Mirouze (2013) Weaver, A. T. and Mirouze, I. (2013). On the diffusion equation and its application to isotropic and anisotropic correlation modelling in variational assimilation. Quarterly Journal of the Royal Meteorological Society, 139 (670), 242–260. doi:10.1002/qj.1955.
    URL https://doi.org/10.1002/qj.1955
  • Weaver et al. (2016) Weaver, A. T., Tshimanga, J. and Piacentini, A. (2016). Correlation operators based on an implicitly formulated diffusion equation solved with the Chebyshev iteration. Quarterly Journal of the Royal Meteorological Society, 142 (694), 455–471. doi:10.1002/qj.2664.
    URL https://doi.org/10.1002/qj.2664
  • Zuo et al. (2024, in preparation) Zuo, H., Alonso-Balmaseda, M., de Boisseson, E., Browne, P. A., Chrust, M., Keely, S., Mogensen, K., Pelletier, C., de Rosnay, P. and Takakura, T. (2024, in preparation). The 6thsuperscript6𝑡6^{th}6 start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT generation of ecmwf ensemble (re)analysis system for ocean and sea-ice. ECMWF Newsletter,.
  • Zuo et al. (2015) Zuo, H., Balmaseda, M. A. and Mogensen, K. (2015). The ECMWF-MyOcean2 eddy-permitting ocean and sea-ice reanalysis ORAP5. Part 1: Implementation. Technical Memorandum 736, ECMWF, Reading, UK.
    URL https://www.ecmwf.int/en/elibrary/76996-ecmwf-myocean2-eddy-permitting-ocean-and-sea-ice-reanalysis-orap5-part-1
  • Zuo et al. (2019) Zuo, H., Balmaseda, M. A., Tietsche, S., Mogensen, K. and Mayer, M. (2019). The ECMWF operational ensemble reanalysis–analysis system for ocean and sea ice: a description of the system and assessment. Ocean Science, 15 (3), 779–808. doi:10.5194/os-15-779-2019.
    URL https://os.copernicus.org/articles/15/779/2019/