Improving Vortex Position Accuracy with a New Multiscale Alignment Ensemble Filter

Yue Ying aNansen Environmental and Remote Sensing Center, Bergen, Norway
bNational Center for Atmospheric Research, Boulder, Colorado

Search for other papers by Yue Ying in
Current site
Google Scholar
PubMed
Close
https://orcid.org/0000-0001-9988-3488
,
Jeffrey L. Anderson bNational Center for Atmospheric Research, Boulder, Colorado

Search for other papers by Jeffrey L. Anderson in
Current site
Google Scholar
PubMed
Close
, and
Laurent Bertino aNansen Environmental and Remote Sensing Center, Bergen, Norway

Search for other papers by Laurent Bertino in
Current site
Google Scholar
PubMed
Close
Open access

Abstract

A multiscale alignment (MSA) ensemble filtering method was introduced by Ying to reduce nonlinear position errors effectively during data assimilation. The MSA method extends the traditional ensemble Kalman filter (EnKF) to update states from large to small scales sequentially, during which it leverages the displacement vectors derived from the large-scale analysis increments to reduce position errors at smaller scales through warping of the model grid. This study stress tests the MSA method in various scenarios using an idealized vortex model. We show that the MSA improves filter performance as number of scales (Ns) increases in the presence of nonlinear position errors. We tuned localization parameters for the cross-scale EnKF updates to find the best performance when assimilating an observation network. To further reduce the scale mismatch between observations and states, a new option called MSA-O is introduced to decompose observations into scale components during assimilation. Cycling DA experiments show that the MSA-O consistently outperforms the traditional EnKF at equal computational cost. A more challenging scenario for the MSA is identified when the large-scale background flow and the small-scale vortex are incoherent in terms of their errors, making the displacement vectors not effective in reducing vortex position errors. Observation availability for the small scales also limits the use of large Ns for the MSA. Potential remedies for these issues are discussed.

© 2023 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Yue Ying, yue.ying@nersc.no

Abstract

A multiscale alignment (MSA) ensemble filtering method was introduced by Ying to reduce nonlinear position errors effectively during data assimilation. The MSA method extends the traditional ensemble Kalman filter (EnKF) to update states from large to small scales sequentially, during which it leverages the displacement vectors derived from the large-scale analysis increments to reduce position errors at smaller scales through warping of the model grid. This study stress tests the MSA method in various scenarios using an idealized vortex model. We show that the MSA improves filter performance as number of scales (Ns) increases in the presence of nonlinear position errors. We tuned localization parameters for the cross-scale EnKF updates to find the best performance when assimilating an observation network. To further reduce the scale mismatch between observations and states, a new option called MSA-O is introduced to decompose observations into scale components during assimilation. Cycling DA experiments show that the MSA-O consistently outperforms the traditional EnKF at equal computational cost. A more challenging scenario for the MSA is identified when the large-scale background flow and the small-scale vortex are incoherent in terms of their errors, making the displacement vectors not effective in reducing vortex position errors. Observation availability for the small scales also limits the use of large Ns for the MSA. Potential remedies for these issues are discussed.

© 2023 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding author: Yue Ying, yue.ying@nersc.no

1. Introduction

Geophysical models have increasing prediction skill over the past decade thanks to improving model resolution, better representation of physical processes, as well as the increasing number and better use of observations that help better initialize the models. Data assimilation (DA) is the process of combining the model forecasts with observational information and finding the best estimate of the model states. DA is an important component in a prediction system, an inferior method can be the bottleneck in prediction skill even though better models and observing networks are already available. The ensemble Kalman filter (EnKF; Evensen 1994; Burgers et al. 1998) is the state-of-the-art DA method in many geophysical prediction problems (Houtekamer and Zhang 2016; Carrassi et al. 2018). However, the EnKF analysis becomes more suboptimal as nonlinearity increases in the system. Hoffman et al. (1995) showed that forecast error can be decomposed into three components: displacement, amplitude, and residual. A common source of nonlinearity is from large position errors (displacement) in geophysical features, such as atmospheric fronts, ocean eddies, sea ice edge, sunspots, and so on. Chen and Snyder (2007) showed in a simple two-dimensional vortex example that a linear error in vortex position manifests as nonlinear errors in the model states (winds), and they suggested including features (position, intensity, and shape) as observations in DA to correct model states. The feature-based observations improve the analyses when position errors are moderate. However, as position errors increase, more nonlinearity is introduced in the relation between observed features and model states, making the EnKF updates suboptimal.

To address this issue, previous studies used two different strategies to improve DA performance. The first strategy is to use a nonlinear DA method, such as the particle filter (PF; Doucet et al. 2001; van Leeuwen 2009; Poterjoy 2016) and rank histogram filter (Anderson 2010, 2019), to obtain the best analysis in the presence of high nonlinearity. Some methods formulate the ensemble filter using non-Gaussian error distributions (moderate position errors will cause a skewed error distributions) (e.g., Bishop 2016; Hodyss et al. 2017; Poterjoy 2022), or transform the variables using Gaussian anamorphosis (Simon and Bertino 2009; Amezcua and van Leeuwen 2014), to address the consequences of nonlinearity. Iterative methods are introduced to find a nonlinear solution with a sequence of linear updates (Sakov et al. 2012; Bocquet and Sakov 2014). These methods can handle nonlinearity in general, not only for position errors but also for nonlinear model dynamics and observation operators, but the drawback is the increased complexity in implementation and additional computational cost compared to the EnKF.

The second strategy for dealing with nonlinear position error is to explicitly account for displacements in DA (Ravela et al. 2007; Beezley and Mandel 2008) or to use the feature alignment technique (FAT; Nehrkorn et al. 2014, 2015; Stratman et al. 2018) prior to DA. For the particular case of hurricane DA, some studies suggested using a storm-relative framework (Aksoy 2013; Navarro and Hakim 2014; Lin et al. 2018) or vortex relocation (Liu et al. 2020) to avoid the negative impact from large displacements. The FAT is a more general solution to the problem, which derives displacement vectors from observations to warp the model grid and reduce the misalignment between the observed and model simulated features, bringing the DA problem closer to a linear regime. Ying (2019) noticed that smoothing a feature to the large scale (low resolution) will naturally reduce the nonlinearity in its position errors and introduced a multiscale alignment (MSA) method that derives displacement vectors more efficiently than the FAT. In the MSA, a model state is decomposed into several scale components (SCs) that are updated by the EnKF sequentially from the large to the small scales. After each update, displacement vectors are derived from the analysis increments and then used to warp the model grid to reduce the position errors at smaller scales (assuming that large-scale and small-scale displacement errors are coherent).

The multiscale approach is proven beneficial in large-dimensional DA problems (Zhang et al. 2009; Miyoshi and Kondo 2013; Buehner and Shlyaeva 2015; Caron and Buehner 2018). Decomposing the model state into SCs allows errors to be represented separately at each scale. Ying (2020) formulated a multiscale approach where both model states and observations are decomposed into SCs, allowing flexibility in the choice of filter parameters (localization, inflation, etc.). The observation SCs are akin to idea of smoothing or differentiating observations to extract information for DA (Weng and Zhang 2012; Bédard and Buehner 2020; Sodhi and Fabry 2022). In this study, we add the decomposition of observation SCs as a new option in the MSA, which is referred to as MSA-O.

The MSA/MSA-O method provides new means of representing the nonlinear error relation across scales through the alignment steps, i.e., using large-scale errors to reduce displacement errors at smaller scales. The existing multiscale DA approaches (e.g., Li et al. 2015; Wang et al. 2021) do not offer this capability since they only model the linear correlation between errors at different scales (off-diagonal terms in the error covariance matrix for SCs). In terms of treatment of position errors, the vortex relocation approach arbitrarily removes position uncertainties based on vortex position observation, while the FAT simultaneously reduces position errors with other errors in minimizing a nonlinear cost function. The MSA is similar to the FAT in finding displacements. However, unlike the FAT, the MSA uses an iteration over scales for better efficiency and to avoid local minima in minimization.

Although the MSA was shown to outperform the EnKF (Ying 2019), several questions still remain about its performance. First, will the MSA analysis be as good as the EnKF analysis in a quasi-linear regime? Second, is the MSA robust, i.e., does it guarantee a performance improvement for all kinds of prior error conditions? Finally, the MSA inevitably brings additional computational and tuning costs as the number of SCs increases. Does the improvement in performance justify the additional cost? In this study, we design numerical experiments using a simplified two-dimensional vortex model, similar to the ones used in Chen and Snyder (2007) and Poterjoy (2022) (the Rankine vortex example in his section 5), to stress-test the MSA method and answer these questions. The model, DA methods, and numerical experiments are described in section 2. Results from idealized experiments are presented in section 3. Cycling DA experiment results are presented in section 4 to provide some insight for future implementation and a discussion for potential issues in more realistic scenarios. A summary of our findings is given in section 5.

2. Methodology

a. Model state and truth

The vortex model, detailed in appendix A, describes a nondivergent horizontal wind field, v = (u, υ), defined on a doubly periodic uniform grid r = (x, y). The initial wind field consists of a Rankine vortex and a background steering flow:
v(r)=vvort(r;Vmax,Rmw,a)+vbkg(r;V¯bkg),
where Vmax, Rmw, and a are the radius of maximum wind, maximum wind speed, and shape parameter for the Rankine vortex, respectively; and V¯bkg is the average of the initial random κ−3 background wind speed.

Model physics is based on two-dimensional vorticity dynamics, but we added a vorticity generation term to increase vortex intensity when the maximum wind speed is lower than a critical speed limit, simulating the rapid intensification process. Despite its simplicity, the vortex model captures the key physical processes and scale interactions in vortex dynamics, and the rapid intensification is particularly interesting and challenging from a DA perspective.

Let ψ be the state vector containing all model variables in physical space, i.e., v(r) flattened into a vector, the model advances from time t to t + 1 as
ψt+1=mtt+1(ψt;ξ,κg,σκ,Vc,ν),
where m is the nonlinear model, ξ is the vorticity generation coefficient, κg and σκ control the spectral shape of the generation term, Vc is the critical speed limit, and ν is the dissipation coefficient.

For the truth (denoted with a star) ψ*, the vortex is placed at the center of the domain, Vmax = 35 m s−1, Rmw = 45 km, and a = 1.5, while the background flow V¯bkg=5ms1. Model parameters are set as follows: the dissipation ν = 50 m2 s−1, the generation term ξ = 8.3 × 10−7 s−1, κg = 8, σκ = 3, and Vc = 75 m s−1. With this setup, the simulated vortex intensifies and reaches the speed limit in ∼9 h, then its intensity oscillates near this limit. However, adaptive time steps are required to achieve steady simulation of this oscillation (otherwise numerical instability occurs). For the sake of simplicity we choose to only run the model for 12 h and focus on DA experiments during the rapid intensification period.

b. Ensemble forecasts

An ensemble of model runs characterize the forecast errors prior to DA. Let n = 1, …, N index the ensemble members. The initial ensemble ψn,t=0 is generated through perturbing the truth in several different ways to represent different sources of uncertainties. Figure 1 gives an example of vortex evolution in an ensemble forecast.

Fig. 1.
Fig. 1.

Vortex evolution from a 20-member ensemble forecast with initial vortex position error Lsprd/Rmw = 0.2, no structure error, out-of-phase background flow errors, and using the imperfect models. Vortex (a) tracks and (b) intensities are shown (colored lines for members, black line for the truth). (c) The domain-averaged model state error (solid black line) and the averaged ensemble spread (dotted line).

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

For member n, its initial wind field is
vn(r)=vvort*(r+rn;Vmax,n,Rmw,n)+vbkg*(r+rbkg,n)+vn.
Here, rn=(xn,yn) is a spatially constant vector field that displaces the wind field to represent position uncertainties, where xn and yn are randomly drawn from a normal distribution N(0,Lsprd). The position spread Lsprd normalized by Rmw is a good indicator of the degree of nonlinearity in wind errors due to displacement, Lsprd/Rmw = 0.2 is a quasi-linear regime and Lsprd/Rmw = 1 enters a highly nonlinear regime.
Vortex structure uncertainties can be accounted for by randomly perturbing the Vmax and Rmw parameter for each member as
Vmax,n=Vmax*+Vn,
Rmw,n=Rmw*+Rn,
where VnN(0,Vsprd) and RnN(0,Rsprd).

The background flow is perturbed either in phase or out of phase. In the in-phase scenario the background flow is displaced along with the vortex (rbkg,n=rn; in phase with the vortex position errors) and an additional additive noise is included to account for other error sources in the wind field (vn). The additive noise is generated in the same manner as the background flow but with a lower magnitude than the background flow V¯=1.5ms1. In the scenario where the background flow errors are out of phase with the vortex errors, rbkg,n=0 and V¯ is increased to 3 m s−1 (in this case the coherence assumption is broken).

Model errors can be introduced by perturbing the vorticity generation coefficient ξ. In a perfect-model scenario, the forecast model is the same as the true model (ξn=ξ*=8.3×107s1). For imperfect models, ξn is randomly drawn from a uniform distribution U(3.3,10)×107s1, this range of values is low-biased to mimic the behavior of real hurricane models that tend to underestimate vortex intensity rather than to overestimate.

c. Observations

Let ϕ be a vector containing the observations, and No be the number of observed locations in the domain. Synthetic observations are generated from the truth as ϕo=h(ψ*)+εo, where h is the observation operator and εoN(0,σo2I) is the observation error. The observation errors are uncorrelated (diagonal covariance), and σo is the observation error standard deviation.

Several different observation scenarios are considered in this study (Table 1). First, in section 3a, we assimilate an observation at one single location (randomly placed in the domain) to study its impact and the asymptotic behavior of the MSA method. Here, in the Single_Wind_Obs scenario, the observation measures directly the u and υ wind at a given location. In this case, h consists mostly of linear interpolation coefficients, and we set σo = 3 m s−1. Then, in the Position_Obs scenario [section 3a(2)] we consider a feature-based observation of the vortex center position (xc, yc). The h operator in this case is nonlinear, involving a search for the maximum in the vorticity field. For position observations, we set σo = 0.1Rmw = 4.5 km. In section 3b, we assimilate several observations from a network. Two different network geometries are tested: first, a global network covering the entire domain (No = 1000 randomly placed locations, ∼8% of which are within a 180-km radius of the vortex center); second, a more realistic targeted network that only has sparse observations within the vicinity of the vortex (No = 60 randomly placed locations, all within a 180-km radius of the vortex center). Finally, in section 4, we conduct cycling DA experiments, assimilating the targeted observation networks (following the true vortex) that are available every 3 h during the 12-h period (in 3 analysis cycles).

Table 1.

List of scenarios and their corresponding configurations in prior ensemble (spread in vortex position, intensity and size, background flow types), assimilated observation type, and forecast model. Multiple values are listed for a parameter if they are tested in a sensitivity experiment. For the cycling experiments (section 4), the vortex and background flow configurations in the table are for the initial conditions.

Table 1.

d. The DA method and experiments

DA finds the optimal combination of the prior ensemble ψnb with the observations ϕo, so that the analysis ensemble ψna has minimum error while its spread matches this error. Here we drop subscript t for simplicity since the analysis takes place at the same time (filtering). The EnKF first computes observation priors (observed values based on prior model states):
ϕnb=h(ψnb),
then compares it with the observations to obtain an innovation ϕoϕnb, which is used to update the state variables through the error covariance:
cov(ψb,ϕb)=1N1n=1N(ψnbψb)(ϕnbϕb)T.
The underlying ideas for the MSA method are decomposing the model states into SCs and applying the EnKF updates separately and utilizing the large-scale analysis increments to derive displacement vectors that reduces position errors at smaller scales.

Appendix B describes methods for scale decomposition. Let Ns be the number of SCs. For scale s, the prior state becomes ψn,sb=Fsψn, where Fs projects the model state onto SCs. This prior state is updated by the innovation through a cross-scale covariance cov(ψsb,ϕb). We propose a new option for the MSA, which we call MSA-O (the letter O stands for observation), where the observations are also decomposed into SCs ϕso=Fsoϕo and used in correspondence with the state SCs, with covariances cov(ψsb,ϕsb), to reduce the scale mismatch during the EnKF updates. Figure 2 illustrates the scale decomposition of u wind states and observations for the Ns = 4 case.

Fig. 2.
Fig. 2.

Illustration of scale decomposition of model state and observation for Ns = 4: (a) kinetic energy spectra (u2+υ2)/2 for the total state (black line) and for each SC (colored lines), (b) u wind state, (c) observed u wind from the targeted network, (d)–(g) u wind state SCs for s = 1–4, and (h)–(k) observation SCs for s = 1–4.

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

In this study, we choose the ensemble square root filter (EnSRF; Whitaker and Hamill 2002; Tippett et al. 2003) to perform the EnKF update. We refer to the MSA method using Ns SCs as EnSRF_MSA_Ns, the MSA-O method as EnSRF_MSA-O_Ns. Note that Ns = 1 reverts to the original EnSRF, and the largest Ns is 7, which is appropriate given the model grid resolution. Appendix C provides a detailed formulation of the MSA/MSA-O method.

To provide a benchmark for comparison, we run free model forecasts from the prior ensemble without assimilation, which is called the NoDA case. For single observation assimilation, a simple PF analysis based on importance sampling (see section 3 in van Leeuwen 2015) is computed to provide another benchmark. The PF applies Bayes rule directly. The prior weight for each member is 1/N, which is multiplied by observation likelihood exp[(ϕoϕnb)2/(2σo2)], then normalized across members to get the posterior weight. The probability distribution is then divided into N equal parts to obtain the analysis ensemble members with equal weights again, in which members with large weights are duplicated while those with small weights are removed. The PF analysis converges to the optimal solution as the ensemble size increases. However, it is not feasible for an observing network with large No since with computationally viable ensemble sizes its weights quickly collapse causing filter degeneracy.

Table 1 summarizes all the test scenarios, in which the different filter methods are compared, and the next section describes the error metrics we used.

e. Evaluation metrics

The first error metric is the domain-averaged root-mean-square state variable error (domain-averaged error for short):
εdomain-avg=[(ψψ*)T(ψψ*)¯]1/2,
where angle brackets 〈·〉 denote an average over the ensemble members and the overbar ()¯ denotes an average over the state variables (over the domain).

Three feature-based metrics help further characterize errors in simulated vortex features: position, intensity, and size. The vortex position is found by searching for the maximum vorticity in the domain. The vortex intensity is defined as the maximum wind speed. The vortex size Rsize characterizes the area of large wind speeds, it is derived from the azimuthal-average wind profile V(R) relative to the diagnosed vortex center, so that V(RmwRRsize) > 15 m s−1 (this value is arbitrarily chosen, for hurricanes the threshold is usually higher to relate to destructive winds). These diagnostic features are commonly used in the hurricane community (e.g., Zhang et al. 2009). Note that the size feature is not the same as Rmw, since it also depends on Vmax (intensity feature).

Let hfeature denote the operator from model state to one of the diagnostic features. Due to its nonlinearity, the mean of features from members does not equal the feature from the ensemble mean. We thus define ensemble-averaged feature errors as
εfeature=[hfeature(ψ)hfeature(ψ*)]21/2.
In sections 3a and 3b, we present errors diagnosed from the posterior ensemble (final analysis) for each DA method. In section 4 (cycling DA) we present both the errors from the posterior ensemble at the analysis cycles and the ensemble forecasts to the end of DA period.

3. Idealized experiment results

a. Assimilation of observations at a single location

1) Asymptotic behavior of the MSA method

We first test how the MSA method performs as Ns increases using the Single_Wind_Obs scenario. Since there is only one location observed in space, it is not possible to decompose observations into SCs (MSA-O is not tested here). We test the EnSRF_MSA with Ns = 1–7.

We first run the experiment once using a large ensemble size N = 200 without localization and visualize the filter updates in physical space in a moderate-nonlinearity scenario (Lsprd/Rmw = 0.6). Figure 3 compares wind contours from the analysis ensemble (members with different colors). The original EnSRF analysis suffers most from nonlinearity, vortices with larger position errors become deformed after the linear updates. On the other hand, the PF analysis perfectly preserves vortex structure in each member by design. The EnSRF_MSA analyses improve as Ns increases in that the vortices are less deformed as they are nudged closer to the truth. There is no guarantee that the MSA method preserves vortex structure and there is still a slight deformation even for Ns = 7.

Fig. 3.
Fig. 3.

Ensemble spaghetti plots of wind speed contours at 20 m s−1 for (a) NoDA, (b) EnSRF (Ns = 1), (c)–(e) EnSRF_MSA (Ns = 3, 5, and 7), and (f) PF for the Single_Wind_Obs scenario with Lsprd/Rmw = 0.6. The + sign marks the observation location and × marks the state variable location for Fig. 5. The first 20 members are plotted in color and the true contour in black.

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

Figure 4 shows wind speed maps from one ensemble member that is displaced to the southeast in the prior. The EnSRF assimilates the observation at + and inserted a new vortex near the true position, but the linear update is not enough to completely remove the prior vortex, causing the analysis vortex to be elongated. A special case EnSRF_MSnoA_5, where the MSA is applied without the alignment steps (similar to multiscale DA approaches), shows that multiscale DA alone does not improve the result. EnSRF_MSA_5 shows that displacement steps greatly contribute to the improvement of vortex structure. Figures 4e–h illustrate the displacement vectors derived from the EnSRF update at s = 1–4, the warping of the grid gradually nudges the vortex toward truth, improving its position accuracy but at a slight cost of distortion of structure.

Fig. 4.
Fig. 4.

Vortex from member 1 compared for (a) NoDA, (b) EnSRF, (c) multiscale EnSRF without alignment (EnSRf_MSnoA_5), and (d) EnSRF_MSA_5. (e)–(h) Intermediate stages after scale iteration s = 1–4 during EnSRF_MSA_5. Wind speed (m s−1) is shown in gray shadings, the 20 m s−1 contour is highlighted in red, and the true contour in black. The observation is located at the + sign. Displacement vectors derived for each scale are shown in (e)–(h).

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

Figure 5 shows the bivariate relations between the observation (located at the + sign in Fig. 3) and one state variable (located at the × sign in Fig. 3) during DA. The EnSRF makes linear updates so the assimilation of u and υ observations shows as two sets of parallel line segments, which tend to push members away from the main cluster in the prior distribution. Although the final analysis ensemble mean gets much closer to the truth, some members have unphysical values in wind that correspond to the deformation of vortices. The PF update line segments are not parallel and the analysis members converge near the true values. The MSA method behaves more like the nonlinear PF update as Ns increases, the members are moved closer to the truth through the sequential updates over the SCs.

Fig. 5.
Fig. 5.

Scatterplots showing the bivariate relation between the observed (a)–(c) u and (d)–(f) υ variable with a u state variable being updated (in this case y axes show the predictor and x axes show the predictand). The position of the observation is located at the + sign in Fig. 3 and the state variable at the × sign in Fig. 3. Three methods are shown: (left) EnSRF, (center) EnSRF_MSA_5, and (right) PF. Gray lines mark the true values and red lines mark the observed values. Green circles are the prior ensemble and blue triangles are the analysis ensemble. Line segments are plotted to show the intermediate stages for 5 members: for the EnSRF case, one segment for assimilating u, then another segment for assimilating υ; for the EnSRF_MSA_5 case, one segment for each scale iteration s = 1, …, 5; for the PF case, only one segment is shown.

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

Now we test the robustness of this asymptotic behavior with 1000 random realizations (trials) using a smaller ensemble size (N = 20) for the EnSRF and EnSRF_MSA, but the PF benchmark is run with larger ensemble size (N = 500). Localization is not applied when assimilating the single observation. The left column of Fig. 6 summarizes the analysis errors averaged over these realizations.

Fig. 6.
Fig. 6.

Error boxplots showing the interquartile range of error metrics [(a),(b) domain-averaged; (c),(d) vortex position; (e),(f) intensity; and (g),(h) size errors] obtained from 1000 trials. Two scenarios are compared: (left) Single_Wind_Obs and (right) Position_Obs. Cases with varying nonlinearity (Lsprd/Rmw from 0.2 to 1.0) are tested. In each case, results are shown for NoDA (gray), EnSRF (white), EnSRF_MSA_Ns (blue, Ns = 1–7 from left to right), and PF (dark green). Note that NoDA and PF results are hard to see in intensity and size errors since they are very close to zero.

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

For domain-averaged errors (Fig. 6a), the EnSRF_MSA method using larger Ns is better than the EnSRF in nonlinear regimes. As nonlinearity increases, a larger Ns is required to obtain the best analysis. In the quasi-linear regime (Lsprd/Rmw = 0.2), the EnSRF analysis is the best while the EnSRF_MSA analyses are worse. The degradation is due to vortex intensity errors (Fig. 6e) introduced by the alignment steps. The grid warping is a slightly diffusive process due to the use of an Eulerian grid. Since the Rankine vortex wind profile is nondifferentiable at the peak, it is difficult to preserve wind maxima when the displacements are not an integer number of grid spacings and the wind maxima end up in between grid points and get weakened during interpolation. The vortices are weakened by ∼2 m s−1 after warping in EnSRF_MSA (this issue can be resolved by using higher-order interpolation or a Lagrangian mesh to describe model state). When we introduce a little nonlinearity by increasing position spread (Lsprd/Rmw ≤ 0.2), the EnSRF analysis becomes more suboptimal and EnSRF_MSA consistently outperforms the EnSRF. Vortex structures are disrupted in the EnSRF analyses due to the linear updates, causing large intensity and size errors. Note that the EnSRF analysis tends to overestimate the wind maxima while the alignment steps underestimate them.

The PF perfectly preserves vortex features for each member by design, thus its intensity and size errors are close to zero (not exactly zero because of the round-off errors in hfeature). In the limit of N → ∞, the PF analysis ensemble represents the true Bayesian posterior distribution. However, in nonlinear regimes, the PF posterior mean has large variance of position errors and domain-averaged errors, while the EnSRF_MSA with larger Ns seemingly outperforms the PF. This is caused by the multimodality of the true posterior distribution given some observations that are ambiguous in matching the prior wind field (next subsection provides a more detailed discussion). For these observations, the PF posterior mean has near zero likelihood of being the truth since it is in between the modes of the posterior distribution. Even with very large ensemble size (N = 500 is used, we have also tried N = 20 000 that exceeds the state dimension without seeing much improvement), the PF still produces larger domain-averaged errors in the nonlinear regime (Lsprd/Rmw = 1) due to this multimodality. The EnSRF_MSA seemingly produces a more accurate posterior mean than the PF when Ns increases. This is because the larger-scale SCs are less multimodal in terms of observation likelihood and the alignment steps tend to collapse the posterior distribution over the correct mode, though at a risk of overfitting the large-scale observation information.

2) Sensitivity to observation type and location

A single wind observation with ±3 m s−1 error is less informative than a position observation with an error of ±4.5 km. Especially when the wind observation is placed farther away from the vortex center, it only provides limited information about the vortex itself. To compare to the Single_Wind_Obs scenario, we use the Position_Obs scenario to test the direct assimilation of vortex position observations and results are shown in the right column of Fig. 6.

In nonlinear regimes, EnSRF_MSA_7 reaches the best performance in the Position_Obs scenario no matter how much nonlinearity arises from different Lsprd, while its performance degrades as Lsprd increases in the Single_Wind_Obs scenario. This shows that the position observations have more impact. However, assimilating position observations is quite challenging using the EnSRF (or when Ns is small for EnSRF_MSA). Because the mapping between the observation space (position) and the state space (winds) is highly nonlinear, the EnSRF is suboptimal and causes distortion of the vortex structures. These distortions are even worse (larger intensity and size errors) in the Position_Obs scenario compared to the Single_Wind_Obs scenario, in agreement with Chen and Snyder (2007) who documented vorticity distortion after assimilating position observations.

The information content of wind observations depends on their locations relative to the true vortex center Robs. Observations near Rmw have the highest observed wind speeds (∼40 m s−1) and with the fixed instrument error of ±3 m s−1 these observations have lower percentage errors (we did not consider the case of real wind observations where observation errors increase with wind speeds). Figure 7 shows the domain-averaged errors (from Fig. 6a) again but with the 1000 trials further categorized into five bins with respect to observation location Robs/Rmw. In the linear regime (Lsprd/Rmw = 0.2), observations farther away (Robs/Rmw > 1.2) cause the degradation in EnSRF_MSA analyses. In the nonlinear regime (Lsprd/Rmw = 1), all observations reduce errors; however, the biggest improvements come from assimilating observations near the maximum wind (RobsRmw) and near the vortex center (Robs ≈ 0). The PF posterior mean has large errors when observations are near the vortex center (Robs/Rmw < 0.8) with low wind speeds. Because both observations from the center and farther outside have low wind speeds, weights are assigned to both members that match the low wind speeds, causing the posterior distribution to be multimodal (ambiguous observation likelihood). The posterior mean is arguably not a good indicator for the PF performance, since it has low probability of being the truth when the true distribution is multimodal. On the other hand, the EnSRF_MSA does collapse the posterior ensemble over the correct mode even though the true Bayesian posterior is multimodal. The larger-scale SCs are less ambiguous when matching with the observations. The alignment produces a more accurate posterior mean, but at a risk of overfitting the large-scale information from the observations, since the true Bayesian posterior is supposed to be multimodal.

Fig. 7.
Fig. 7.

Boxplots for domain-averaged wind errors (as in Fig. 6a) but summarized individually with respect to position spread (Lsprd/Rmw, increasing from left to right) and vicinity of wind observation to the true vortex center (Robs/Rmw, increasing from bottom up). With the five categories, there are about 200 trials in each.

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

b. Assimilation of observation networks

1) Tuning for best multiscale localization

In this subsection, we perform manual tuning to select the localization parameters for a multiscale framework. Two parameters are tuned through empirical evaluation: the radius of influence (ROI) and the amplitude factor α (see appendix C for definition). The tested values are ROI = 8, 12, 16, 20, 24, 28, 32, 40, 48, 64 grid points and α = 0, 0.1, 0.2, …, 1, which results in 110 cases in total. For each case, 100 trials are run where the global network is assimilated to update a prior ensemble (N = 20 members) with position spread Lsprd/Rmw = 0.6 and out-of-phase background flow errors.

The filter update for only the first SC in EnSRF_MSA_Ns is diagnosed for domain-averaged errors. Figure 8a shows the range of parameters that yields good performance (best performance marked as a star). As Ns increases, the first SC corresponds to larger scales and there is a bigger scale mismatch between the observations and the states. The best ROI increases with the underlying correlation scale, which is expected. On the other hand, the best α decreases as Ns increases, and the reasons are twofold. Because the bigger scale mismatch introduces more sampling noise that is not necessarily distance-dependent, reducing α helps reduce the impact of the noise. Also, as a larger ROI is favored for the larger scale, more observations are participating in the update, reducing α helps to prevent overfitting the observations.

Fig. 8.
Fig. 8.

Tuning of the localization ROI (in units of number of Δx) and amplitude parameter (α) for updating the first SC using (a) the MSA and (b) the MSA-O algorithms. Domain-averaged wind errors are obtained from 100 trials for each combination of parameters and tested for cases with Ns from 2 to 7 (color-coded). The best performance (minimum error) for each case is marked with a star, the contours indicate the range of parameters that achieve an error within 1% of the best performance.

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

Localization for the MSA-O algorithm is tuned in the same manner and results are shown in Fig. 8b. In this case, since observations are also decomposed to SCs, the scale mismatch is much reduced in the update step, which requires less α adjustment (i.e., α is closer to 1). However, the best performance still comes from α < 1 as in the MSA case, only a much wider range of parameters give relatively good performance (cf. the purple lines, Ns = 7). The MSA-O algorithm requires smaller α than the MSA when updating smaller scales (cf. red lines, Ns = 2), because the difference between the irregular observation network and the uniform model grid (see Fig. 2) further introduces discrepancies between the observation space and state space. Based on these findings, we configure the MSA in the following experiments with the best-performing localization parameters, which are listed in Table 2. The MSA-O configuration (not listed) is very similar to the MSA, only a few differences for small Ns.

Table 2.

Tuned localization parameters (ROIs and αs for s = 1, …, Ns from left to right) for the MSA method.

Table 2.

We do not expect the best localization parameters for scale s + 1 to depend on those for scale s (hence no need for iterative tuning). The best parameters for a range of characteristic scales have been identified in our tuning for s = 1, Ns = 1–7. We use these results to guide the choice for s = 2–Ns listed in Table 2. For example, in Ns = 7 the s = 2 setting comes from the Ns = 5; s = 1 result, since they are at about the same spatial scale. Although they are not exactly the same SC (only matching characteristic wavenumber), we believe this approach is sufficient, given that the performance is not that sensitive to exact choice of parameters (see wide range of good performance in parameter space in Fig. 8).

2) Sensitivity to uncertainties other than position

We created scenarios with different sources of initial uncertainties, in background flow (Incoherent_BkgFlow) and in vortex structure (Vmax_Error and Rmw_Error), to test the robustness of the MSA performance. Figure 9 compares the EnSRF_MSA performance under these scenarios.

Fig. 9.
Fig. 9.

Comparison of error boxplots (from 100 trials) among NoDA (gray), EnSRF (white), and EnSRF_MSA_Ns (blue; Ns = 2–6 shown from left to right). Four scenarios with different initial condition uncertainties are tested (a),(e),(i),(m) Baseline; (b),(f),(j),(n) Incoherent_BkgFlow; (c),(g),(k),(o) Vmax_Error; and (d),(h),(l),(p) Rmw_Error (see Table 1 for details). Four error metrics are shown from the top to bottom rows.

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

The MSA method is built on the assumption that small-scale features and large-scale patterns are coherent so that the large-scale updates can be utilized to correct for errors in the small-scale features. However, it is debatable whether this assumption holds for all geophysical modeling scenarios. In Incoherent_BkgFlow, we test how the MSA method behaves when deviating from the coherent assumption. The incoherent background flow errors (see section 2b for details) are challenging for the MSA. Notice in Fig. 9f how the EnSRF_MSA analysis becomes worse than the EnSRF analysis as Ns increase (some improvement can be seen as Ns keep increasing, but still worse than Ns = 1). In the large-scale SC, there are both vortex-scale features projected to the large scale and the large-scale background flow uncertainties themselves. In Incoherent_BkgFlow, the alignment step can sometimes displace the vortex erroneously according to analysis increments after correcting the additive noise in the background flow, which increases the vortex position errors and degrades the analysis. In Baseline, the MSA method works as expected with best performance.

In Vmax_Error and Rmw_Error, the uncertainties in vortex structure are accounted for by varying the two parameters during generation of the initial Rankine vortex. Figures 9g and 9h show that the additional uncertainties do not impact the performance of the MSA as the position errors are reduced as expected when Ns increases. Perturbing Vmax increases both vortex intensity and size errors in the analysis (size also depends on Vmax due to its definition). Perturbing Rmw increases vortex size errors in the analysis.

For all scenarios, the intensity errors first decrease then increase when Ns increases from 1 to 6 (Figs. 9i–l). EnSRF (Ns = 1) tends to overestimate the vortex intensity with occasional overfitting to local observations of peak wind speeds. On the other hand, EnSRF_MSA with large Ns tends to underestimate the vortex intensity because of the alignment and warping of the grid weakens the vortex [see discussion in section 3a(1)]. Some degradation in vortex size can be seen for larger Ns (Figs. 9o,p). See Figs. 4e–h how the vortex becomes more asymmetric as more warping steps are applied. As a network of many observations is assimilated here, the smaller-scale analysis increments and the displacement vectors introduce even more distortion to the wind field (compared to the single observation results in Fig. 4).

3) Sensitivity to observation density and coverage

This subsection tests the sensitivity of the MSA/MSA-O performance to different observation density and coverage (i.e., its information content). Two types of observing network are considered: the global network is denser with coverage of the entire domain (No = 1000), while the targeted network is more sparse (No = 60) and is only deployed near the vortex center (within 180 km radius, see Fig. 2c for a visualization). Figure 10 compares the performance of EnSRF_MSA and EnSRF_MSA-O (Ns = 1–4) when assimilating these two types of observations.

Fig. 10.
Fig. 10.

Comparison of error boxplots (from 100 trials) among NoDA (gray), EnSRF (white), EnSRF_MSA (blue; Ns = 2–4 shown from left to right), and EnSRF_MSA-O (red; Ns = 2–4 shown from left to right). Two scenarios are tested: (left) assimilating global network (Baseline) and (right) Targeted_Network. For each scenario, three cases with Lsprd/Rmw = 0.2, 0.6, and 1.0 are tested. Four error metrics are shown from top to bottom.

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

Let us first look at the quasi-linear regime (Lsprd/Rmw = 0.2) in Global_Network. Decomposition of observation SCs in MSA-O improves position accuracy compared to MSA (Fig. 10c) but at a cost of degrading the domain-averaged error (Fig. 10a). Comparing the state SCs (Figs. 2d–g) with the observation SCs (Figs. 2h–k), we see that the smaller scales have significant mismatches due to differences in grid geometry. When No is not large enough, aliasing may occur. Although the MSA-O benefits from assimilating the large-scale observation SCs with better match to the state SCs, the inaccuracies in the scale decomposition (mostly at small scales, κ > 8) bring negative impacts. In real implementations, the smaller-scale observation SCs can be discarded (i.e., not assimilated) to avoid the negative impact if its accuracy is low. As Ns increases, the MSA/MSA-O also tends to weaken the vortex intensity (Fig. 10e) due to warping of the grid [see section 3a(1)]. The size accuracy (Fig. 10g) does not monotonically improve as Ns increases, larger Ns degrades the size feature, probably because the MSA/MSA-O overfits the small-scale pattern emerging from the observations that are contaminated by aliasing noises. In the nonlinear regime (Lsprd/Rmw = 1), the behavior of MSA/MSA-O is the same. Since higher nonlinearity occurs in the prior errors the MSA/MSA-O consistently outperforms the single-scale EnSRF.

Using Global_Network as a baseline, now we compare the performance in Targeted_Network. Notice that the global network constrains the domain-averaged errors much better than the targeted network, thanks to its better information content. For Targeted_Network in the nonlinear regime (Fig. 10b), the benefit from using observation SCs (MSA-O) out-weights the negative impact from observation SC inaccuracies. Position rrors (Fig. 10d) show that MSA-O performs better than MSA when assimilating the targeted network. Because the targeted network has smaller-scale vortex information dominating the larger-scale background flow information, the observation SCs help to extract information about the large scales in MSA-O, which helps to reduce position errors after assimilation.

4. Cycling data assimilation experiments

In this section, we test the MSA/MSA-O algorithms in a cycling DA setting to provide some context for real application scenarios such as hurricane forecasts. Four scenarios are tested: Baseline, Incoherent_BkgFlow, Vort_Struct_Error, and Imperfect_Model (see Table 1). For each scenario, the targeted observing network is assimilated every 3 h during a 12-h period, the cycling DA is performed for 100 trials over which the error metrics are evaluated from both the analysis ensemble and the ensemble forecasts to the end of the period.

One important constraint we place on the MSA/MSA-O algorithms is that a smaller ensemble size is used for larger Ns so that all methods are compared at equal computational cost. Ensemble size is 30 for NoDA and EnSRF, and 28, 25, and 22 for Ns = 2, 3, and 4. A larger Ns effectively increases the state vector size (see Ying 2019) that is offset by reducing the ensemble size. The extra cost in the alignment step is also balanced by the reduced number of model runs in the ensemble forecast step.

Figure 11 compares the analysis and forecast error boxplots from different methods for each test scenario, and Fig. 12 shows the corresponding error time series, averaged over 100 trials. As discussed in section 3b(3), the MSA-O approach is better than the MSA when assimilating the targeted network. The errors are consistently smaller from EnSRF_MSA-O than EnSRF_MSA, and the EnSRF_MSA-O_3 method produces the best overall performance.

Fig. 11.
Fig. 11.

Cycling DA experiments: comparison of error boxplots (from 100 trials) among NoDA (gray), EnSRF (white), EnSRF_MSA (blue), and EnSRF_MSA-O (red). Four scenarios are tested: (a),(e),(i),(m) Baseline; (b),(f),(j),(n) Incoherent_BkgFlow; (c),(g),(k),(o) Vort_Struct_Error; and (d),(h),(l),(p) Imperfect_Model. Four error metrics are shown (from top to bottom) for the analysis (averaged over cycles at t = 3, 6, and 9 h) and forecast (averaged over the three forecasts at t = 12 h).

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

Fig. 12.
Fig. 12.

As in Fig. 11, but showing averaged error time series from the 100 trials. In each panel, errors from NoDA (gray), EnSRF (black), EnSRF_MSA_3 (blue), and EnSRF_MSA-O_3 (red) are compared. Analysis cycles take place at t = 3, 6, and 9 h. Forecasts are run from the analysis time to the end of the period (t = 12 h) and marked as dotted lines with corresponding colors.

Citation: Monthly Weather Review 151, 6; 10.1175/MWR-D-22-0140.1

In the Baseline scenario, the analysis domain-averaged errors are smaller in EnSRF_MSA-O than in EnSRF, but the difference are small among different Ns, because the targeted network only updates near the vortex, which is a small portion of the entire domain. Clearer differences are seen for the feature errors: MSA-O with increasing Ns further reduces the position and intensity errors, but the improvement for size errors stops at Ns = 3 and starts increasing back for Ns = 4. For the forecasts, EnSRF_MSA-O_3 is the best-performing method. EnSRF_MSA-O_4 is struggling at vortex position and size in the forecasts. The error time series suggest that error grows more rapidly after the EnSRF analyses (black lines) than the EnSRF_MSA-O_3 analyses (red lines), especially for the vortex position errors (Fig. 12e). Although the EnSRF analysis is already very accurate according to the domain-averaged errors, some physical imbalances (large local wind maxima that are not axisymmetric in the vortex) are introduced when overfitting local observations and overestimating intensity (see increased intensity error in Fig. 12i). The more rapid forecast error growth also indicates such imbalances. On the other hand, the MSA-O method is designed to better preserve the vortex physical balance while reducing position errors at the same time. The drawbacks from the MSA-O using larger Ns > 3 is that it tends to overfit the large-scale SCs and collapse ensemble spread in position prematurely (not shown), so that when assimilating the small-scale SCs there is not enough prior spread to allow sufficient increments. The smaller-scale observation SCs becomes more contaminated with aliasing noises (see Figs. 1h–k), the vortex maximum wind features in s = 4 already has very low signal-to-noise ratio, which is why increasing Ns beyond 3 does not bring additional benefits to outweigh the side effects of warping. For Ns = 2 and 3, however, the MSA-O method clearly improved the smaller-scale vortex features through better utilization of the observed information.

In the Incoherent_BkgFlow scenario, the larger initial uncertainties in background flow cause more rapid position error growth and overall larger vortex size errors than the baseline. NoDA position errors exceed 100 km (off chart in Fig. 11f). For the analysis, the EnSRF has the best performance. However, EnSRF_MSA-O_2 gives the overall best forecasts so it has an advantage over the EnSRF (Fig. 11). The deviation from the coherent assumption [see section 3b(2)] is also among the reasons why larger Ns does not further improve the MSA-O results. The behavior of the EnKF having better analysis but worse forecasts is also reported by other studies (e.g., Poterjoy et al. 2017; Poterjoy 2022). Inspection of the EnSRF ensemble spread (not shown) indicates that its forecast spread is also larger and its analysis spread is smaller than those from the EnSRF_MSA-O. This confirms that the EnSRF tends to overfit the observations, causing physical imbalances and more rapid error growth in the forecasts.

The Vort_Struct_Error scenario does not differ too much from the Baseline. With additional uncertainties in the initial vortex structure, the intensity and size errors (Figs. 11k,o) increase for NoDA. But the EnSRF is still able to reduce errors, even a bit more for the intensity and size errors thanks to the increased initial ensemble spread in vortex features. The MSA/MSA-O behave as expected in further improving over the EnSRF.

For the Imperfect_Model scenario, the additional model uncertainties cause much larger errors in vortex intensity than the Baseline, while the position and size errors remain low. NoDA has intensity error off the chart in Fig. 11l. Despite this additional challenge, EnSRF_MSA-O_3 is still superior to the EnSRF for all error metrics except the vortex intensity. This result suggests that the alignment step works as long as patterns exist for the feature of interest (vortex structure in this case). When there are additional amplitude errors in the patterns being aligned, the alignment algorithm still works robustly. These amplitude errors caused by model uncertainties, however, are independent of the position errors, they lead to higher vortex intensity errors, for which the MSA/MSA-O does not improve over the EnSRF.

5. Conclusions and discussion

In this study, we stress test the MSA method of Ying (2019) in a simplified two-dimensional vortex model similar to that of Chen and Snyder (2007), which allows easier creation of various test scenarios and separation of different error sources (vortex position errors, structure errors, background flow errors and model errors). The asymptotic behavior of the MSA method is first demonstrated. As Ns increases, the EnSRF_MSA analysis errors decrease, approaching what can be achieved by a particle filter. We show that the best Ns depends on the degree of nonlinearity caused by initial position spread Lsprd. As position errors increase a larger Ns is required to achieve the best performance. The observation type (u, υ wind, or position observations) and the location of the observation relative to the vortex center (Robs) change the information content for a single observation. The EnSRF_MSA method is less sensitive to observation location and can extract information from feature-based observations better than the original EnSRF. The correction of vortex location will precondition the small-scale features in the prior so that observations of these features can be better assimilated. However, we also revealed several processes that limit the robustness of the benefits (incoherence across scales, aliasing noises in small scales, and weakening of vortex due to linear interpolation).

A coherence assumption is made in the MSA method to reduce position errors in small scales according to large-scale analysis increments. A deviation from the coherence assumption is shown to degrade the performance of the MSA method. When background flow errors are independent (out of phase) of the vortex position errors, the interference from these background flow errors lead to misalignment of the vortices. It is unclear how frequently this incoherence between large-scale and small-scale features occurs in real geophysical prediction problems. To remedy this issue, we suggest further incorporating an adjustment term in the displacement vectors to scale their impact on the smaller scale. The adjustment term can be diagnosed from the ensemble-estimated cross-scale covariance (off-diagonal parts of the multiscale covariance; Fig. 6 in Buehner and Shlyaeva 2015). Incoherence between large- and small-scale components will result in lower cross-scale covariance, then the adjustment term will taper the displacement vectors accordingly. Another option worth trying is to first decompose the model state into individual components that are incoherent with each other, then perform the MSA analysis separately for each. In general, disentangling the independent components and updating them separately seems a good strategy in DA to achieve better performance for each component.

When assimilating a network of observations, the additional decomposition of observation SCs is introduced as a new option, MSA-O, which is shown to perform better than MSA, especially when assimilating limited observation information (the targeted network), thanks to the reduced scale mismatch between observations and state variables during the filter update. The extraction of observation SCs (could be stated as an optimization problem itself) is not necessarily easy and is prone to aliasing errors, especially when observations are sparse and only with partial coverage. However, the benefits from assimilating observation SCs instead of the raw observation (e.g., Bédard and Buehner 2020; Sodhi and Fabry 2022) often outweigh these drawbacks. In our study, the MSA-O method has consistently better performance than the original EnSRF in all the test scenarios.

While this study provides further proof of concept of the MSA method and its feasibility, we do expect some more challenges in real applications. Model biases can be much more severe in real models; features of interest can be completely missing in the forecasts compared to the observations. The MSA method is not expected to work in such extremely nonlinear scenarios. On the other hand, in the quasi-linear regime, the EnSRF method is already the best algorithm so the MSA method cannot improve. Thus, in practice a test for nonlinearity (e.g., Poterjoy and Anderson 2016; Kurosawa and Poterjoy 2023) may be necessary to determine which DA method to use in the analysis. When applying the MSA method, the best choice for Ns may depend on how much nonlinearity there is in the system, it may also depend on the observing networks (sparse observations may prevent the use of large Ns due to lack of information at small scales). Nonlinearity and observation availability may vary across the domain, for example, an atmospheric model might have most of its domain in quasi-linear regimes and only nonlinear regimes along the sharp gradients in frontal zones. For such scenarios, the MSA can be applied in a local analysis, i.e., the frontal zone can be isolated in a subdomain within which the DA analysis takes place. Similar to the idea of nested domains with different resolution, the MSA method can be applied to child domains where high nonlinearity occurs, and not the parent domain where the EnKF is already optimal in the linear regime.

Arguably, the simple vortex model results can be too optimistic since the model does not capture the intricate inner-core dynamics of a real three-dimensional vortex, especially how its asymmetric features interact with its environmental conditions, which are the current forecast challenges. Complications can arise in the implementation in a full-physics multivariable model, some treatment of topography and lower boundary conditions in deriving displacement vectors can be found in Nehrkorn et al. (2014). The results from this study only provide some initial guidance and should not be interpreted generally for all model systems. Future studies shall also investigate the possibility to extend the two-dimensional displacement vector fields to three- or even four-dimensional (include the time dimension to reduce timing errors).

We pay extra attention to computational cost when comparing the MSA with the original EnSRF method. Our results suggest that—at equal cost—the MSA-O algorithm consistently outperforms the EnSRF in nonlinear regimes, suggesting that it is scalable to larger problems. In practice, the cost of tuning filter parameters, such as localization and inflation, is also nonnegligible. The manual tuning of localization parameters in this study would be infeasible in realistic models. The use of adaptive algorithms may thus be needed to provide online estimation of the best parameters, without the need for a priori tuning. Based on the localization literature (e.g., Ying et al. 2018, and references therein), the best localization distance depends on the physical correlation scale of the dynamical system, the ensemble size, and the observation density. For localization, adaptive algorithms (Zhen and Zhang 2014; Moosavi et al. 2018; Cheng et al. 2021) and empirical localization functions (e.g., Anderson and Lei 2013; Lei and Anderson 2014) can be used to automate the tuning. We did not apply any covariance inflation in our study for the sake of simplicity, but for numerous cycles and more complex models, we expect inflation to be necessary. Adaptive algorithms are also available (e.g., El Gharamti 2018) to automate the tuning of inflation. For alignment techniques, inflation methods based on location uncertainties can also be considered (e.g., Zhen et al. 2022). In realistic applications, the truth is unknown and errors are mixed from different sources, which pose challenges to the tuning of DA algorithms. Findings from this study should provide some initial guidance for future implementation of the MSA method in more challenging scenarios.

Acknowledgments.

Yue Ying thanks the NCAR Advanced Study Program for supporting this study in its early stages, this study is also supported by the institutional basic funding from NERSC (Research Council of Norway Project 218857). We appreciate the constructive comments from Jonathan Poterjoy and two anonymous reviewers that improved an earlier version of this manuscript.

Data availability statement.

A Python code package is written to perform all the experiments shown in this study. The code is open-source and available at https://doi.org/10.5281/zenodo.6477080. The package includes the two-dimensional vortex model, data assimilation methods (ensemble filters and the multiscale algorithms), and the scripts that run all the experiments and generate all the figures in this study. The experiment data can be easily reproduced by running the Python code. Error diagnostics data for reproducing the figures are archived at https://doi.org/10.5281/zenodo.7020672.

APPENDIX A

The Two-Dimensional Vortex Model

The model state can be defined by a nondivergent horizontal wind field, (u, υ), on a square domain with periodic boundary conditions and no rotation. The model can be nondimensionalized and applied to vortex systems at different scales, but in this study we formulated the model with physical units commonly found in hurricane prediction to provide some context. The model grid uses Cartesian coordinates r = (x, y) with length L in each direction, it has 128 × 128 grid points and the grid spacing is set to Δx = 9 km. The governing equation is
ζt=vζ+γζ+ν2ζ,
where ζ = ∂υ/∂x − ∂u/∂y is the vorticity, γ is the generation rate, and ν is a dissipation coefficient.
The equation is solved in spectral space using the fourth-order Runge–Kutta numerical scheme with a time step of 60 s. Let κ = (kx, ky) denote the wavenumber in (x, y) directions that form the spectral space, κ = |κ| is the total wavenumber. The generation rate is spatially varying and defined in spectral space as
γ^(κ)={ξexp[(κκg)22σκ2]ifmax(V)<Vc0ifmax(V)Vc,
where γ^ is the Fourier coefficients of γ, ξ is a scalar generation coefficient, Vc is the critical wind speed that limits the generation, κg is the center wavenumber and σκ is the spectral bandwidth of the generation rate. The generation rate has a Gaussian spectral shape so that generation occurs mostly at wavenumbers κg ± σκ.
The wind field consists of a Rankine vortex embedded in a background flow that both evolve with time. The initial Rankine vortex is created so that its center position is rc = (xc, yc) and its radial wind profile follows:
Vvort(R)={VmaxRRmwifRRmwVmax(RmwR)aifR>Rmw,
where Vvort = |vvort|, R = |rrc| is the radius relative to the center, Vmax is the vortex maximum wind speed, Rmw is the radius of maximum wind, and a is a shape parameter.

The background flow is initialized at t = 0 as a random wind field with a spectrum V^bkg(κ)κ3 and wind speed Vbkg = |vbkg|. The average wind speed V¯bkg (the bar denotes a spatial average) can be specified to control the relative strength of the initial background flow, for example V¯bkg=0 turns off the background flow. To generate the random wind field, we first draw a random vorticity field ζ^bkg(κ)κ1, convert the vorticity to wind so that the wind is nondivergent and has the correct −3 power law, and then normalize and scale the wind speed to V¯bkg.

APPENDIX B

The Scale Decomposition Method

Let Ns be the number of SCs. Given L = 128Δx, the largest meaningful wavenumber is κmax = 16 in this study (8Δx resolves a complete sine wave). The characteristic wavenumber for each SC can be defined as κs=κmaxs/Ns, for s = 1, …, Ns. For the sth SC, a spectral lowpass response function is defined as
f^sL(κ)={1if|κ|<κs[cos(|κ|κsκs+1κs)]2ifκs|κ|κs+10if|κ|>κs+1,
and a bandpass response function is then constructed by
f^s={f^sLifs=1f^sLf^s1Lif1<s<Ns1f^s1Lifs=Ns.
In vector form, the spectral-space response function is f^s=f^s(κ), corresponding to fs = fs(r) in physical space. Since model states are defined on a uniform grid, it is easy to transform the states to spectral space, apply the response function in an element-wise product ψ^s=f^sψ^, then transform back to get the sth state SC ψs.
Observation SCs are more difficult to derive because of the irregular grid. According to the convolution theorem, an element-wise product f^sψ^ in spectral space is equivalent to a convolution fsψ in physical space. In this study, the convolution is done through a series of spatially weighted averages. The sth observation SC can be obtained by ϕso=fsϕo, whose ith element is
ϕs,io=j=1Nofs(rirj)ϕjo,
where ri is the position of ϕio. The low-pass filter can also be implemented more efficiently using a diffusion-based algorithm (e.g., Grooms et al. 2021) for the irregular grids. Following Buehner and Shlyaeva (2015), we allow some spectral overlapping between SCs by using a squared cosine transitioning from 1 to 0 in the response functions. This overlapping makes the SC computation more numerically stable, it also allows stronger error correlation among SCs which is more suitable for the MSA.

Hereafter we express the scale decomposition operation in matrix form as ψn,s = Fsψn for the model states, ϕso=Fsoϕo for the observations, and ϕn,sb=Fsoh(ψnb) for the observation priors. Each row of Fs contains the averaging kernel fs, and Fso only differs from Fs due to different geometry of the model grid and the observing network.

Spatial inhomogeneity in the observation grid will result in inaccuracies in the derived SCs. One can pad the data voids with fake observations (zeros; will not be assimilated) to achieve a relatively even distribution. Of course, the same padding needs to be applied to the observation priors (simulated observations from model states) to derive a consistent innovation.

We use a reduced-dimension grid to represent large-scale SCs, which improves the efficiency of the multiscale approaches. The necessary grid dimension decreases with κs. The EnKF update is performed on the coarse grid and increments are then refined (with bilinear interpolation) and added to the native model grid, which is a common approach for multiscale/multigrid DA and in incremental 4DVar (Courtier et al. 1994).

APPENDIX C

The Multiscale Alignment (MSA) Method

The EnKF update equation [Eqs. (13) and (14) in Burgers et al. (1998)] can be restated as
ψna=ψnb+Lcov(ψb,ϕb)cov(ϕb,ϕb)+R(ϕoϕnb),
for n = 1, …, N, where ψnb is the prior state, ϕo is the observation, ϕnb=h(ψnb) is the observation prior, ensemble-estimated error covariances (cov) are computed according to (7), R is the specified observation error covariance, and L is the localization to remove spurious error covariances. The localization factor for assimilating the ith observation and updating the jth state variable is
Li,j=α×GC(|rirj|,ROI),
where GC is the Gaspari and Cohn (1999) localization function with a parameter called radius of influence (ROI; the distance at which the covariance is tapered to zero), α is an additional amplitude parameter (similar to the one used in Lei and Whitaker 2017).

Algorithm 1 The MSA (if decompose_obs = False) and MSA-O (if decompose_obs = True) methods. Note that subscript n implies that the operation is repeated for n = 1, …, N.

Input: prior ensemble ψn, observations ϕo

Output: iteratively updated ψn as final analysis ensemble

Parameters: number of scales Ns, localization function Ls(ROIs, αs), observation error σo,s2, and smoothness constraint w

  • 1: for s in 1, …, Ns do

  • 2:  ψn,sb=Fsψn

  • 3:  ϕnb=h(ψn)

  • 4:  if decompose_obs then

  • 5:    ϕso=Fsoϕo

  • 6:    ϕn,sb=Fsoh(ψn)

  • 7:    ψn,sa=ψn,sb+Lscov(ψsb,ϕsb)cov(ϕsb,ϕsb)+σo,s2I(ϕsoϕn,sb)

  • 8:  else

  • 9:    ψn,sa=ψn,sb+Lscov(ψsb,ϕb)cov(ϕb,ϕb)+σo2I(ϕoϕnb)

  • 10:  end if

  • 11:  if s < Ns then

  • 12:    qn,s=argminqψn,sb(q)ψn,sa2+wq2

  • 13:    ψnψn(qn,s)+ψn,saψn,sb(qn,s)

  • 14:  else

  • 15:    ψnψn+ψn,saψn,sb

  • 16:  end if

  • 17: end for

The MSA method (Ying 2019) applies the EnKF update (C1) sequentially for the SCs. Algorithm 1 provides a pseudocode description. Note that each EnKF update makes an “analysis increment” ψn,saψn,sb, while all the iterations lead to the “final analysis” ψn.

We keep a general form of an EnKF update in the algorithm, indicating that different EnKF variants can be used. In this study, we choose to use the ensemble square root filter variant (EnSRF; Whitaker and Hamill 2002; Tippett et al. 2003), which assimilates observations one at a time assuming their errors to be uncorrelated R=σo2I. We only consider the scenario when σo is specified correctly during DA, and performed some manual tuning to optimize the localization parameters.

After obtaining the analysis increment, displacement vectors q are derived by minimizing the cost function ψb(q)ψa2+wq2 using the Horn and Schunck (1981) optical flow algorithm. The term ψ(q) is a shorthand to represent the warped model states, u(r + q) and υ(r + q). Note that q is a spatially varying vector field, unlike the constant position perturbations rn in (3). The smoothness constraint is set to w = 1 [not tuned, see Horn and Schunck (1981) for a discussion of its effect]. The displacement vectors are then applied to warp the model state to reduce position errors at smaller scales, the warping is done through bilinear interpolation to evaluate ψ(q).

We proposed and tested a new option to decompose observations into SCs corresponding to the state SCs, which we call MSA-O (decompose_obs = True in Algorithm 1). This option allows observations to be closer to the model states (less scale mismatch) during assimilation, improving the filter performance. Note that, for the MSA-O, the observation errors also undergo scale decomposition (B3) so that the observation error standard deviation for ϕso is adjusted to
σo,s=σofs(0),
i.e., large-scale observation error is reduced.

REFERENCES

  • Aksoy, A., 2013: Storm-relative observations in tropical cyclone data assimilation with an ensemble Kalman filter. Mon. Wea. Rev., 141, 506522, https://doi.org/10.1175/MWR-D-12-00094.1.

    • Search Google Scholar
    • Export Citation
  • Amezcua, J., and P. J. van Leeuwen, 2014: Gaussian anamorphosis in the analysis step of the EnKF: A joint state-variable/observation approach. Tellus, 66A, 23493, https://doi.org/10.3402/tellusa.v66.23493.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2010: A non-Gaussian ensemble filter update for data assimilation. Mon. Wea. Rev., 138, 41864198, https://doi.org/10.1175/2010MWR3253.1.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2019: A nonlinear rank regression method for ensemble Kalman filter data assimilation. Mon. Wea. Rev., 147, 28472860, https://doi.org/10.1175/MWR-D-18-0448.1.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., and L. Lei, 2013: Empirical localization of observation impact in ensemble Kalman filters. Mon. Wea. Rev., 141, 41404153, https://doi.org/10.1175/MWR-D-12-00330.1.

    • Search Google Scholar
    • Export Citation
  • Bédard, J., and M. Buehner, 2020: A practical assimilation approach to extract smaller-scale information from observations with spatially correlated errors: An idealized study. Quart. J. Roy. Meteor. Soc., 146, 468482, https://doi.org/10.1002/qj.3687.

    • Search Google Scholar
    • Export Citation
  • Beezley, J. D., and J. Mandel, 2008: Morphing ensemble Kalman filters. Tellus, 60A, 131140, https://doi.org/10.1111/j.1600-0870.2007.00275.x.

    • Search Google Scholar
    • Export Citation
  • Bishop, C. H., 2016: The GIGG-EnKF: Ensemble Kalman filtering for highly skewed non-negative uncertainty distributions. Quart. J. Roy. Meteor. Soc., 142, 13951412, https://doi.org/10.1002/qj.2742.

    • Search Google Scholar
    • Export Citation
  • Bocquet, M., and P. Sakov, 2014: An iterative ensemble Kalman smoother. Quart. J. Roy. Meteor. Soc., 140, 15211535, https://doi.org/10.1002/qj.2236.

    • Search Google Scholar
    • Export Citation
  • Buehner, M., and A. Shlyaeva, 2015: Scale-dependent background-error covariance localisation. Tellus, 67A, 28027, https://doi.org/10.3402/tellusa.v67.28027.

    • Search Google Scholar
    • Export Citation
  • Burgers, G., P. J. Van Leeuwen, and G. Evensen, 1998: Analysis scheme in the ensemble Kalman filter. Mon. Wea. Rev., 126, 17191724, https://doi.org/10.1175/1520-0493(1998)126<1719:ASITEK>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Caron, J.-F., and M. Buehner, 2018: Scale-dependent background error covariance localization: Evaluation in a global deterministic weather forecasting system. Mon. Wea. Rev., 146, 13671381, https://doi.org/10.1175/MWR-D-17-0369.1.

    • Search Google Scholar
    • Export Citation
  • Carrassi, A., M. Bocquet, L. Bertino, and G. Evensen, 2018: Data assimilation in the geosciences: An overview of methods, issues, and perspectives. Wiley Interdiscip. Rev.: Climate Change, 9, e535, https://doi.org/10.1002/wcc.535.

    • Search Google Scholar
    • Export Citation
  • Chen, Y., and C. Snyder, 2007: Assimilating vortex position with an ensemble Kalman filter. Mon. Wea. Rev., 135, 18281845, https://doi.org/10.1175/MWR3351.1.

    • Search Google Scholar
    • Export Citation
  • Cheng, S., J.-P. Argaud, B. Iooss, A. Ponçot, and D. Lucor, 2021: A graph clustering approach to localization for adaptive covariance tuning in data assimilation based on state-observation mapping. Math. Geosci., 53, 17511780, https://doi.org/10.1007/s11004-021-09951-z.

    • Search Google Scholar
    • Export Citation
  • Courtier, P., J.-N. Thépaut, and A. Hollingsworth, 1994: A strategy for operational implementation of 4D-Var, using an incremental approach. Quart. J. Roy. Meteor. Soc., 120, 13671387, https://doi.org/10.1007/s11004-021-09951-z.

    • Search Google Scholar
    • Export Citation
  • Doucet, A., D. F. Nando, and N. Gordon, 2001: Sequential Monte Carlo Methods in Practice. Information Science and Statistics, Vol. 1, Springer, 581 pp.

  • El Gharamti, M., 2018: Enhanced adaptive inflation algorithm for ensemble filters. Mon. Wea. Rev., 146, 623640, https://doi.org/10.1175/MWR-D-17-0187.1.

    • Search Google Scholar
    • Export Citation
  • Evensen, G., 1994: Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics. J. Geophys. Res., 99, 10 14310 162, https://doi.org/10.1029/94JC00572.

    • Search Google Scholar
    • Export Citation
  • Gaspari, G., and S. E. Cohn, 1999: Construction of correlation functions in two and three dimensions. Quart. J. Roy. Meteor. Soc., 125, 723757, https://doi.org/10.1002/qj.49712555417.

    • Search Google Scholar
    • Export Citation
  • Grooms, I., N. Loose, R. Abernathey, J. M. Steinberg, S. D. Bachman, G. Marques, A. P. Guillaumin, and E. Yankovsky, 2021: Diffusion-based smoothers for spatial filtering of gridded geophysical data. J. Adv. Model. Earth Syst., 13, e2021MS002552, https://doi.org/10.1029/2021MS002552.

    • Search Google Scholar
    • Export Citation
  • Hodyss, D., J. L. Anderson, N. Collins, W. F. Campbell, and P. A. Reinecke, 2017: Quadratic polynomial regression using serial observation processing: Implementation within DART. Mon. Wea. Rev., 145, 44674479, https://doi.org/10.1175/MWR-D-17-0089.1.

    • Search Google Scholar
    • Export Citation
  • Hoffman, R. N., Z. Liu, J.-F. Louis, and C. Grassoti, 1995: Distortion representation of forecast errors. Mon. Wea. Rev., 123, 27582770, https://doi.org/10.1175/1520-0493(1995)123<2758:DROFE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Horn, B. K. P., and B. G. Schunck, 1981: Determining optical flow. Artif. Intell., 17, 185203, https://doi.org/10.1016/0004-3702(81)90024-2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., and F. Zhang, 2016: Review of the ensemble Kalman filter for atmospheric data assimilation. Mon. Wea. Rev., 144, 44894532, https://doi.org/10.1175/MWR-D-15-0440.1.

    • Search Google Scholar
    • Export Citation
  • Kurosawa, K., and J. Poterjoy, 2023: A statistical hypothesis testing strategy for adaptively blending particle filters and ensemble Kalman filters for data assimilation. Mon. Wea. Rev., 151, 105125, https://doi.org/10.1175/MWR-D-22-0108.1.

    • Search Google Scholar
    • Export Citation
  • Lei, L., and J. L. Anderson, 2014: Empirical localization of observations for serial ensemble Kalman filter data assimilation in an atmospheric general circulation model. Mon. Wea. Rev., 142, 18351851, https://doi.org/10.1175/MWR-D-13-00288.1.

    • Search Google Scholar
    • Export Citation
  • Lei, L., and J. S. Whitaker, 2017: Evaluating the trade-offs between ensemble size and ensemble resolution in an ensemble-variational data assimilation system. J. Adv. Model. Earth Syst., 9, 781789, https://doi.org/10.1002/2016MS000864.

    • Search Google Scholar
    • Export Citation
  • Li, Z., J. C. McWilliams, K. Ide, and J. D. Farrara, 2015: A multiscale variational data assimilation scheme: Formulation and illustration. Mon. Wea. Rev., 143, 38043822, https://doi.org/10.1175/MWR-D-14-00384.1.

    • Search Google Scholar
    • Export Citation
  • Lin, K.-J., S.-C. Yang, and S. S. Chen, 2018: Reducing TC position uncertainty in an ensemble data assimilation and prediction system: A case study of Typhoon Fanapi (2010). Wea. Forecasting, 33, 561582, https://doi.org/10.1175/WAF-D-17-0152.1.

    • Search Google Scholar
    • Export Citation
  • Liu, Q., and Coauthors, 2020: Vortex initialization in the NCEP operational hurricane models. Atmosphere, 11, 968, https://doi.org/10.3390/atmos11090968.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., and K. Kondo, 2013: A multi-scale localization approach to an ensemble Kalman filter. SOLA, 9, 170173, https://doi.org/10.2151/sola.2013-038.

    • Search Google Scholar
    • Export Citation
  • Moosavi, A., A. Attia, and A. Sandu, 2018: A machine learning approach to adaptive covariance localization. arXiv, 1801.00548v3, https://doi.org/10.48550/arXiv.1801.00548.

  • Navarro, E. L., and G. J. Hakim, 2014: Storm-centered ensemble data assimilation for tropical cyclones. Mon. Wea. Rev., 142, 23092320, https://doi.org/10.1175/MWR-D-13-00099.1.

    • Search Google Scholar
    • Export Citation
  • Nehrkorn, T., B. Woods, T. Auligné, and R. N. Hoffman, 2014: Application of feature calibration and alignment to high-resolution analysis: Examples using observations sensitive to cloud and water vapor. Mon. Wea. Rev., 142, 686702, https://doi.org/10.1175/MWR-D-13-00164.1.

    • Search Google Scholar
    • Export Citation
  • Nehrkorn, T., B. K. Woods, R. N. Hoffman, and T. Auligné, 2015: Correcting for position errors in variational data assimilation. Mon. Wea. Rev., 143, 13681381, https://doi.org/10.1175/MWR-D-14-00127.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., 2016: A localized particle filter for high-dimensional nonlinear systems. Mon. Wea. Rev., 144, 5976, https://doi.org/10.1175/MWR-D-15-0163.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., 2022: Implications of multivariate non-Gaussian data assimilation for multi-scale weather prediction. Mon. Wea. Rev., 150, 14751493, https://doi.org/10.1175/MWR-D-21-0228.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., and J. L. Anderson, 2016: Efficient assimilation of simulated observations in a high-dimensional geophysical system using a localized particle filter. Mon. Wea. Rev., 144, 20072020, https://doi.org/10.1175/MWR-D-15-0322.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., R. A. Sobash, and J. L. Anderson, 2017: Convective-scale data assimilation for the Weather Research and Forecasting Model using the local particle filter. Mon. Wea. Rev., 145, 18971918, https://doi.org/10.1175/MWR-D-16-0298.1.

    • Search Google Scholar
    • Export Citation
  • Ravela, S., K. Emanuel, and D. McLaughlin, 2007: Data assimilation by field alignment. Physica D, 230, 127145, https://doi.org/10.1016/j.physd.2006.09.035.

    • Search Google Scholar
    • Export Citation
  • Sakov, P., D. S. Oliver, and L. Bertino, 2012: An iterative EnKF for strongly nonlinear systems. Mon. Wea. Rev., 140, 19882004, https://doi.org/10.1175/MWR-D-11-00176.1.

    • Search Google Scholar
    • Export Citation
  • Simon, E., and L. Bertino, 2009: Application of the Gaussian anamorphosis to assimilation in a 3-D coupled physical-ecosystem model of the North Atlantic with the EnKF: A twin experiment. Ocean Sci., 5, 495510, https://doi.org/10.5194/os-5-495-2009.

    • Search Google Scholar
    • Export Citation
  • Sodhi, J. S., and F. Fabry, 2022: Benefits of smoothing backgrounds and radar reflectivity observations for multiscale data assimilation with an ensemble Kalman filter at convective scales: A proof-of-concept study. Mon. Wea. Rev., 150, 589601, https://doi.org/10.1175/MWR-D-21-0130.1.

    • Search Google Scholar
    • Export Citation
  • Stratman, D. R., C. K. Potvin, and L. J. Wicker, 2018: Correcting storm displacement errors in ensembles using the feature alignment technique (FAT). Mon. Wea. Rev., 146, 21252145, https://doi.org/10.1175/MWR-D-17-0357.1.

    • Search Google Scholar
    • Export Citation
  • Tippett, M. K., J. L. Anderson, C. H. Bishop, T. M. Hamill, and J. S. Whitaker, 2003: Ensemble square root filters. Mon. Wea. Rev., 131, 14851490, https://doi.org/10.1175/1520-0493(2003)131<1485:ESRF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • van Leeuwen, P. J., 2009: Particle filtering in geophysical systems. Mon. Wea. Rev., 137, 40894114, https://doi.org/10.1175/2009MWR2835.1.

    • Search Google Scholar
    • Export Citation
  • van Leeuwen, P. J., 2015: Nonlinear data assimilation for high-dimensional systems. Nonlinear Data Assimilation, P. J. van Leeuwen, Y. Cheng, and S. Rich, Eds., Springer, 1–73, https://doi.org/10.1007/978-3-319-18347-3_1.

  • Wang, X., H. G. Chipilski, C. H. Bishop, E. Satterfield, N. Baker, and J. S. Whitaker, 2021: A multiscale local gain form ensemble transform Kalman filter (MLGETKF). Mon. Wea. Rev., 149, 605622, https://doi.org/10.1175/MWR-D-20-0290.1.

    • Search Google Scholar
    • Export Citation
  • Weng, Y., and F. Zhang, 2012: Assimilating airborne Doppler radar observations with an ensemble Kalman filter for convection-permitting hurricane initialization and prediction: Katrina (2005). Mon. Wea. Rev., 140, 841859, https://doi.org/10.1175/2011MWR3602.1.

    • Search Google Scholar
    • Export Citation
  • Whitaker, J. S., and T. M. Hamill, 2002: Ensemble data assimilation without perturbed observations. Mon. Wea. Rev., 130, 19131924, https://doi.org/10.1175/1520-0493(2002)130<1913:EDAWPO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Ying, Y., 2019: A multiscale alignment method for ensemble filtering with displacement errors. Mon. Wea. Rev., 147, 45534565, https://doi.org/10.1175/MWR-D-19-0170.1.

    • Search Google Scholar
    • Export Citation
  • Ying, Y., 2020: Assimilating observations with spatially correlated errors using a serial ensemble filter with a multiscale approach. Mon. Wea. Rev., 148, 33973412, https://doi.org/10.1175/MWR-D-19-0387.1.

    • Search Google Scholar
    • Export Citation
  • Ying, Y., F. Zhang, and J. L. Anderson, 2018: On the selection of localization radius in ensemble filtering for multiscale quasigeostrophic dynamics. Mon. Wea. Rev., 146, 543560, https://doi.org/10.1175/MWR-D-17-0336.1.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., Y. Weng, J. A. Sippel, Z. Meng, and C. H. Bishop, 2009: Cloud-resolving hurricane initialization and prediction through assimilation of Doppler radar observations with an ensemble Kalman filter. Mon. Wea. Rev., 137, 21052125, https://doi.org/10.1175/2009MWR2645.1.

    • Search Google Scholar
    • Export Citation
  • Zhen, Y., and F. Zhang, 2014: A probabilistic approach to adaptive covariance localization for serial ensemble square root filters. Mon. Wea. Rev., 142, 44994518, https://doi.org/10.1175/MWR-D-13-00390.1.

    • Search Google Scholar
    • Export Citation
  • Zhen, Y., V. Resseguier, and B. Chapron, 2022: Physically constrained covariance inflation from location uncertainty and optimal transportation. arXiv, 2211.04207v2, https://doi.org/10.48550/arXiv.2211.04207.

Save
  • Aksoy, A., 2013: Storm-relative observations in tropical cyclone data assimilation with an ensemble Kalman filter. Mon. Wea. Rev., 141, 506522, https://doi.org/10.1175/MWR-D-12-00094.1.

    • Search Google Scholar
    • Export Citation
  • Amezcua, J., and P. J. van Leeuwen, 2014: Gaussian anamorphosis in the analysis step of the EnKF: A joint state-variable/observation approach. Tellus, 66A, 23493, https://doi.org/10.3402/tellusa.v66.23493.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2010: A non-Gaussian ensemble filter update for data assimilation. Mon. Wea. Rev., 138, 41864198, https://doi.org/10.1175/2010MWR3253.1.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2019: A nonlinear rank regression method for ensemble Kalman filter data assimilation. Mon. Wea. Rev., 147, 28472860, https://doi.org/10.1175/MWR-D-18-0448.1.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., and L. Lei, 2013: Empirical localization of observation impact in ensemble Kalman filters. Mon. Wea. Rev., 141, 41404153, https://doi.org/10.1175/MWR-D-12-00330.1.

    • Search Google Scholar
    • Export Citation
  • Bédard, J., and M. Buehner, 2020: A practical assimilation approach to extract smaller-scale information from observations with spatially correlated errors: An idealized study. Quart. J. Roy. Meteor. Soc., 146, 468482, https://doi.org/10.1002/qj.3687.

    • Search Google Scholar
    • Export Citation
  • Beezley, J. D., and J. Mandel, 2008: Morphing ensemble Kalman filters. Tellus, 60A, 131140, https://doi.org/10.1111/j.1600-0870.2007.00275.x.

    • Search Google Scholar
    • Export Citation
  • Bishop, C. H., 2016: The GIGG-EnKF: Ensemble Kalman filtering for highly skewed non-negative uncertainty distributions. Quart. J. Roy. Meteor. Soc., 142, 13951412, https://doi.org/10.1002/qj.2742.

    • Search Google Scholar
    • Export Citation
  • Bocquet, M., and P. Sakov, 2014: An iterative ensemble Kalman smoother. Quart. J. Roy. Meteor. Soc., 140, 15211535, https://doi.org/10.1002/qj.2236.

    • Search Google Scholar
    • Export Citation
  • Buehner, M., and A. Shlyaeva, 2015: Scale-dependent background-error covariance localisation. Tellus, 67A, 28027, https://doi.org/10.3402/tellusa.v67.28027.

    • Search Google Scholar
    • Export Citation
  • Burgers, G., P. J. Van Leeuwen, and G. Evensen, 1998: Analysis scheme in the ensemble Kalman filter. Mon. Wea. Rev., 126, 17191724, https://doi.org/10.1175/1520-0493(1998)126<1719:ASITEK>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Caron, J.-F., and M. Buehner, 2018: Scale-dependent background error covariance localization: Evaluation in a global deterministic weather forecasting system. Mon. Wea. Rev., 146, 13671381, https://doi.org/10.1175/MWR-D-17-0369.1.

    • Search Google Scholar
    • Export Citation
  • Carrassi, A., M. Bocquet, L. Bertino, and G. Evensen, 2018: Data assimilation in the geosciences: An overview of methods, issues, and perspectives. Wiley Interdiscip. Rev.: Climate Change, 9, e535, https://doi.org/10.1002/wcc.535.

    • Search Google Scholar
    • Export Citation
  • Chen, Y., and C. Snyder, 2007: Assimilating vortex position with an ensemble Kalman filter. Mon. Wea. Rev., 135, 18281845, https://doi.org/10.1175/MWR3351.1.

    • Search Google Scholar
    • Export Citation
  • Cheng, S., J.-P. Argaud, B. Iooss, A. Ponçot, and D. Lucor, 2021: A graph clustering approach to localization for adaptive covariance tuning in data assimilation based on state-observation mapping. Math. Geosci., 53, 17511780, https://doi.org/10.1007/s11004-021-09951-z.

    • Search Google Scholar
    • Export Citation
  • Courtier, P., J.-N. Thépaut, and A. Hollingsworth, 1994: A strategy for operational implementation of 4D-Var, using an incremental approach. Quart. J. Roy. Meteor. Soc., 120, 13671387, https://doi.org/10.1007/s11004-021-09951-z.

    • Search Google Scholar
    • Export Citation
  • Doucet, A., D. F. Nando, and N. Gordon, 2001: Sequential Monte Carlo Methods in Practice. Information Science and Statistics, Vol. 1, Springer, 581 pp.

  • El Gharamti, M., 2018: Enhanced adaptive inflation algorithm for ensemble filters. Mon. Wea. Rev., 146, 623640, https://doi.org/10.1175/MWR-D-17-0187.1.

    • Search Google Scholar
    • Export Citation
  • Evensen, G., 1994: Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics. J. Geophys. Res., 99, 10 14310 162, https://doi.org/10.1029/94JC00572.

    • Search Google Scholar
    • Export Citation
  • Gaspari, G., and S. E. Cohn, 1999: Construction of correlation functions in two and three dimensions. Quart. J. Roy. Meteor. Soc., 125, 723757, https://doi.org/10.1002/qj.49712555417.

    • Search Google Scholar
    • Export Citation
  • Grooms, I., N. Loose, R. Abernathey, J. M. Steinberg, S. D. Bachman, G. Marques, A. P. Guillaumin, and E. Yankovsky, 2021: Diffusion-based smoothers for spatial filtering of gridded geophysical data. J. Adv. Model. Earth Syst., 13, e2021MS002552, https://doi.org/10.1029/2021MS002552.

    • Search Google Scholar
    • Export Citation
  • Hodyss, D., J. L. Anderson, N. Collins, W. F. Campbell, and P. A. Reinecke, 2017: Quadratic polynomial regression using serial observation processing: Implementation within DART. Mon. Wea. Rev., 145, 44674479, https://doi.org/10.1175/MWR-D-17-0089.1.

    • Search Google Scholar
    • Export Citation
  • Hoffman, R. N., Z. Liu, J.-F. Louis, and C. Grassoti, 1995: Distortion representation of forecast errors. Mon. Wea. Rev., 123, 27582770, https://doi.org/10.1175/1520-0493(1995)123<2758:DROFE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Horn, B. K. P., and B. G. Schunck, 1981: Determining optical flow. Artif. Intell., 17, 185203, https://doi.org/10.1016/0004-3702(81)90024-2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., and F. Zhang, 2016: Review of the ensemble Kalman filter for atmospheric data assimilation. Mon. Wea. Rev., 144, 44894532, https://doi.org/10.1175/MWR-D-15-0440.1.

    • Search Google Scholar
    • Export Citation
  • Kurosawa, K., and J. Poterjoy, 2023: A statistical hypothesis testing strategy for adaptively blending particle filters and ensemble Kalman filters for data assimilation. Mon. Wea. Rev., 151, 105125, https://doi.org/10.1175/MWR-D-22-0108.1.

    • Search Google Scholar
    • Export Citation
  • Lei, L., and J. L. Anderson, 2014: Empirical localization of observations for serial ensemble Kalman filter data assimilation in an atmospheric general circulation model. Mon. Wea. Rev., 142, 18351851, https://doi.org/10.1175/MWR-D-13-00288.1.

    • Search Google Scholar
    • Export Citation
  • Lei, L., and J. S. Whitaker, 2017: Evaluating the trade-offs between ensemble size and ensemble resolution in an ensemble-variational data assimilation system. J. Adv. Model. Earth Syst., 9, 781789, https://doi.org/10.1002/2016MS000864.

    • Search Google Scholar
    • Export Citation
  • Li, Z., J. C. McWilliams, K. Ide, and J. D. Farrara, 2015: A multiscale variational data assimilation scheme: Formulation and illustration. Mon. Wea. Rev., 143, 38043822, https://doi.org/10.1175/MWR-D-14-00384.1.

    • Search Google Scholar
    • Export Citation
  • Lin, K.-J., S.-C. Yang, and S. S. Chen, 2018: Reducing TC position uncertainty in an ensemble data assimilation and prediction system: A case study of Typhoon Fanapi (2010). Wea. Forecasting, 33, 561582, https://doi.org/10.1175/WAF-D-17-0152.1.

    • Search Google Scholar
    • Export Citation
  • Liu, Q., and Coauthors, 2020: Vortex initialization in the NCEP operational hurricane models. Atmosphere, 11, 968, https://doi.org/10.3390/atmos11090968.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., and K. Kondo, 2013: A multi-scale localization approach to an ensemble Kalman filter. SOLA, 9, 170173, https://doi.org/10.2151/sola.2013-038.

    • Search Google Scholar
    • Export Citation
  • Moosavi, A., A. Attia, and A. Sandu, 2018: A machine learning approach to adaptive covariance localization. arXiv, 1801.00548v3, https://doi.org/10.48550/arXiv.1801.00548.

  • Navarro, E. L., and G. J. Hakim, 2014: Storm-centered ensemble data assimilation for tropical cyclones. Mon. Wea. Rev., 142, 23092320, https://doi.org/10.1175/MWR-D-13-00099.1.

    • Search Google Scholar
    • Export Citation
  • Nehrkorn, T., B. Woods, T. Auligné, and R. N. Hoffman, 2014: Application of feature calibration and alignment to high-resolution analysis: Examples using observations sensitive to cloud and water vapor. Mon. Wea. Rev., 142, 686702, https://doi.org/10.1175/MWR-D-13-00164.1.

    • Search Google Scholar
    • Export Citation
  • Nehrkorn, T., B. K. Woods, R. N. Hoffman, and T. Auligné, 2015: Correcting for position errors in variational data assimilation. Mon. Wea. Rev., 143, 13681381, https://doi.org/10.1175/MWR-D-14-00127.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., 2016: A localized particle filter for high-dimensional nonlinear systems. Mon. Wea. Rev., 144, 5976, https://doi.org/10.1175/MWR-D-15-0163.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., 2022: Implications of multivariate non-Gaussian data assimilation for multi-scale weather prediction. Mon. Wea. Rev., 150, 14751493, https://doi.org/10.1175/MWR-D-21-0228.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., and J. L. Anderson, 2016: Efficient assimilation of simulated observations in a high-dimensional geophysical system using a localized particle filter. Mon. Wea. Rev., 144, 20072020, https://doi.org/10.1175/MWR-D-15-0322.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., R. A. Sobash, and J. L. Anderson, 2017: Convective-scale data assimilation for the Weather Research and Forecasting Model using the local particle filter. Mon. Wea. Rev., 145, 18971918, https://doi.org/10.1175/MWR-D-16-0298.1.

    • Search Google Scholar
    • Export Citation
  • Ravela, S., K. Emanuel, and D. McLaughlin, 2007: Data assimilation by field alignment. Physica D, 230, 127145, https://doi.org/10.1016/j.physd.2006.09.035.

    • Search Google Scholar
    • Export Citation
  • Sakov, P., D. S. Oliver, and L. Bertino, 2012: An iterative EnKF for strongly nonlinear systems. Mon. Wea. Rev., 140, 19882004, https://doi.org/10.1175/MWR-D-11-00176.1.

    • Search Google Scholar
    • Export Citation
  • Simon, E., and L. Bertino, 2009: Application of the Gaussian anamorphosis to assimilation in a 3-D coupled physical-ecosystem model of the North Atlantic with the EnKF: A twin experiment. Ocean Sci., 5, 495510, https://doi.org/10.5194/os-5-495-2009.

    • Search Google Scholar
    • Export Citation
  • Sodhi, J. S., and F. Fabry, 2022: Benefits of smoothing backgrounds and radar reflectivity observations for multiscale data assimilation with an ensemble Kalman filter at convective scales: A proof-of-concept study. Mon. Wea. Rev., 150, 589601, https://doi.org/10.1175/MWR-D-21-0130.1.

    • Search Google Scholar
    • Export Citation
  • Stratman, D. R., C. K. Potvin, and L. J. Wicker, 2018: Correcting storm displacement errors in ensembles using the feature alignment technique (FAT). Mon. Wea. Rev., 146, 21252145, https://doi.org/10.1175/MWR-D-17-0357.1.

    • Search Google Scholar
    • Export Citation
  • Tippett, M. K., J. L. Anderson, C. H. Bishop, T. M. Hamill, and J. S. Whitaker, 2003: Ensemble square root filters. Mon. Wea. Rev., 131, 14851490, https://doi.org/10.1175/1520-0493(2003)131<1485:ESRF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • van Leeuwen, P. J., 2009: Particle filtering in geophysical systems. Mon. Wea. Rev., 137, 40894114, https://doi.org/10.1175/2009MWR2835.1.

    • Search Google Scholar
    • Export Citation
  • van Leeuwen, P. J., 2015: Nonlinear data assimilation for high-dimensional systems. Nonlinear Data Assimilation, P. J. van Leeuwen, Y. Cheng, and S. Rich, Eds., Springer, 1–73, https://doi.org/10.1007/978-3-319-18347-3_1.

  • Wang, X., H. G. Chipilski, C. H. Bishop, E. Satterfield, N. Baker, and J. S. Whitaker, 2021: A multiscale local gain form ensemble transform Kalman filter (MLGETKF). Mon. Wea. Rev., 149, 605622, https://doi.org/10.1175/MWR-D-20-0290.1.