1. Introduction
Variational data assimilation (DA) methods based on tangent linear models (TLMs) continue to show superior forecast accuracy when compared to pure ensemble-based methods or TLM-free hybrid methods (Lorenc et al. 2015; Bowler et al. 2017). However, such TLMs are difficult to maintain and develop, especially for high-resolution models of the coupled Earth system. Local ensemble TLMs (LETLMs) (Frolov and Bishop 2016; Allen et al. 2017; Bishop et al. 2017) can potentially replace traditional TLMs by using a linear matrix operator constructed from an ensemble of nonlinear forward model integrations. A hybrid four-dimensional variational data assimilation (hybrid-4DVAR) algorithm based on an LETLM can combine the benefits of traditional 4DVAR methods, such as propagation of the static covariance through time and four-dimensional localization of the ensemble covariance, with the low development costs and high computational scalability of the TLM-free methods like hybrid-4DEnVAR (Buehner et al. 2010; Kleist and Ide 2015). In fact, our previous work (Allen et al. 2017; Bishop et al. 2017) showed that it is possible to replace a traditional TLM within the context of a hybrid-4DVAR system with no loss in assimilation skill. However, these previous papers developed and tested ensemble-based TLMs only in the context of simplified models: a coupled 1D Lorenz model (Bishop et al. 2017) and a global shallow water model (Allen et al. 2017). To date, there have been no demonstrations that the LETLM can show comparable skill to a traditional TLM in the context of a realistic 3D model of the Earth’s atmosphere.
In this paper, we compare the ability of the traditional TLM and the LETLM to propagate 4DVAR analysis increments generated by an operational weather prediction system run at a degraded resolution. Specifically, we used the Navy Global Environmental Model (NAVGEM) with 2.5° horizontal resolution and 60 vertical levels (surface to about 65 km). We used the degraded resolution to expedite the development of the system. We plan to evaluate the applicability of our findings to a higher-resolution model in future publications.
We compared the ability of the LETLM to propagate an analysis increment against a pair of nonlinear forecasts and against the operational TLM that linearizes the Navy Operational Global Atmospheric Prediction System (NOGAPS)—the Eulerian predecessor to NAVGEM. Our choice of NOGAPS TLM reflects the reality that TLMs used by operational centers often lag the developments in the forecast model. Our tests evaluated the ability of the LETLM to propagate increments over a wide range of altitudes, including the atmospheric boundary layer, troposphere, stratosphere, and lower mesosphere, and also examine the sensitivity of the LETLM to inclusion of radiative forcing and physics parameterizations.
The layout of the paper is as follows. Section 2 describes the NAVGEM, the traditional TLM, and the LETLM. Section 3 details the experimental design, error metrics, and tuning methodology, while the appendix describes the efficient matrix inverse implementation. The experimental results are presented first in section 4, focusing on the tuning of a reference configuration for the LETLM and detailed evaluation of the LETLM performance for a single date. Sensitivity of the LETLM to variations of the model parameters is described in section 5. In section 6, we evaluate the performance of the LETLM for seven different dates, with the goal of establishing statistical significance of the results established for a single date in sections 4 and 5. Section 7 quantifies the computational efficiency and scalability of the LETLM. The summary and conclusions are provided in section 8, and some thoughts on future work are given in section 9.
2. Model description
a. NAVGEM
This study uses the global atmospheric NAVGEM (Hogan et al. 2014), which uses a semi-Lagrangian–semi-implicit integration of the hydrostatic dynamical equations, the first law of thermodynamics, and conservation of moisture and ozone. This study uses a 60-level hybrid sigma-pressure coordinate (top at 0.05 hPa, or about 65 km) as described in Eckermann et al. (2009). To facilitate multiple sensitivity experiments, the model was run at relatively low resolution of T47 (144 longitudes × 72 latitudes, for a Gaussian grid spacing of ~2.5° at the equator). The model time step was 1800 s. The same forecast configuration and resolution was used for the control and the ensemble forecasts. The predicted variables of the NAVGEM are vorticity, divergence, virtual potential temperature, specific humidity Q, and surface pressure Ps. In addition, the zonal U and meridional wind V, temperature T, geopotential height Z,
The NAVGEM code incorporates several stochastic physics packages, including stochastic kinetic energy backscatter (SKEB), nonorographic gravity wave drag, and a stochastic mass flux parameterization in the boundary layer. Stochastic processes are difficult to model from a linear perspective, since there is no deterministic process that results in the random physics change. For the sake of this initial pilot study with NAVGEM, we decided to turn off these stochastic processes for testing the propagation of analysis increments.
The NAVGEM DA solver is a hybrid-4DVAR system (Kuhl et al. 2013) that employs a strong-constraint approach with an ensemble system based on the ensemble transform (ET) method. The cycling window is 6 h, centered at 0000, 0600, 1200, and 1800 UTC. The background includes the last 6 h of a 9-h forecast initialized at the center of the previous analysis time window. The solver employs the accelerated representer (AR) method (Xu et al. 2005; Rosmond and Xu 2006).
b. Traditional tangent linear model
The NAVGEM TLM and adjoint models are described in Rosmond (1997). While the nonlinear forecast model is semi-Lagrangian, the TLM uses a spectral code based on the earlier NOGAPS. This earlier version of the NOGAPS TLM is currently used in the operational 4DVAR system, while a new semi-Lagrangian TLM was still in the development stage at the time of the publication. The TLM is run with the same horizontal and vertical resolution as the nonlinear model, but with a smaller time step of 900 s. The TLM neglects several physical processes, such as gravity wave drag, radiation, and ozone photochemistry. We will refer to this TLM as the NOGAPS TLM in this paper.
c. Local ensemble tangent linear model








In this paper, we focus on a specific form of the ensemble-based TLM—the LETLM—that predicts the future state at each grid point based on the information in a surrounding influence volume. An example of such an influence volume is shown in Fig. 1, where the points in the influence volume with radius 2000 km are shown with red dots, and the location of the predicted point is shown with a green dot located near Chicago (41°N, 90°W). The influence volume also has a vertical dimension (discussed below), which may incorporate multiple model levels above or below the level that is being predicted, resulting in a three-dimensional influence volume in the shape of a cylinder.

An example of an LETLM influence volume (L = 2000 km) on a T47 Gaussian grid. Red dots show the influence region associated with the central point shown in green, located at 41°N, 90°W. There are 200 points in this influence volume.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

An example of an LETLM influence volume (L = 2000 km) on a T47 Gaussian grid. Red dots show the influence region associated with the central point shown in green, located at 41°N, 90°W. There are 200 points in this influence volume.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
An example of an LETLM influence volume (L = 2000 km) on a T47 Gaussian grid. Red dots show the influence region associated with the central point shown in green, located at 41°N, 90°W. There are 200 points in this influence volume.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1



















3. Experiment setup
a. Cycling data assimilation system
For the LETLM experiments described in this paper, we required example analysis increments to propagate with the LETLM and, as explained in section 3b, a database of analysis increments to use as perturbations to initialize the 6-h ensemble forecasts required by the LETLM. The analysis increments for both the test cases and the ensemble perturbations were obtained from a run of the hybrid-4DVAR system of Kuhl et al. (2013). We configured the hybrid-4DVAR system to use only the climatological error covariance model so that it was just a pure 4DVAR formulation. One reason we chose this experiment configuration was that our prior work showed that the ability of the LETLM to propagate analysis corrections depends on the weighting given to the ensemble and climatological parts of the hybrid background error covariance model. The LETLM was found to perform at its worst (best) when full (zero) weight is given to the climatological part of the hybrid matrix (Frolov and Bishop 2016; Allen et al. 2017). Hence, to make our experiments more stringent, we chose to address the harder problem of propagating a climatological perturbation. Another reason was that degrading the system to pure 4DVAR removed any aliasing between the analysis increment and the ensemble. To ensure consistency among the model first guess, the ensemble resolution, the background trajectory for the TLM, and the 4DVAR inner loop, we used the same T47L60 (about 2.5° at the equator, or 277 km) resolution for all components of the cycling system. In our experiments, we assimilated the full suite of conventional observations and Advanced Microwave Sounding Unit (AMSU)-A microwave radiances.
The 4DVAR system was initialized from a fully spun up, high-resolution (T425L60) model state that was degraded to T47L60. We cycled the system for 120 days, starting at 0000 UTC 15 November 2014. The cycling run generated a set of 480 analysis increments, 400 of which were used as perturbations to initialize the ensemble (analysis increments between 21 November 2014 and 2 March 2015). Most of the experiments were performed using a reference configuration that was initialized from the same initial condition at 0600 UTC 20 November 2014. This date was chosen because it allowed for most of the initial shocks from the resolution downgrade to fully dissipate, and it allowed the cycling system to settle into a new steady state. In addition to the reference configuration that was tuned for the cycle at 0600 UTC 20 November 2014, we performed a second tuning based at 0000 UTC 4 March 2015. We then tested the LETLM performance using the average of these two tunings on seven additional experiments from a randomly selected set of times over the period 5–15 March 2015. Figure 2 summarizes our experimental configuration above and highlights the fact that the analysis increments used to generate ensemble perturbations came from a period that was distinct from the period used for testing of the LETLM performance.

Schematic of the experimental period. (a) Spin up and calibration, 15–20 Nov 2014; (b) archive of ensemble perturbations, 20 Nov 2014 to 2 Mar 2015; and (c) LETLM testing, 4–15 Mar 2015.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Schematic of the experimental period. (a) Spin up and calibration, 15–20 Nov 2014; (b) archive of ensemble perturbations, 20 Nov 2014 to 2 Mar 2015; and (c) LETLM testing, 4–15 Mar 2015.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Schematic of the experimental period. (a) Spin up and calibration, 15–20 Nov 2014; (b) archive of ensemble perturbations, 20 Nov 2014 to 2 Mar 2015; and (c) LETLM testing, 4–15 Mar 2015.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
b. Ensemble generation
Computation of the LETLM relies on the availability of ensemble forecasts. These ensemble forecasts can come from a cycling ensemble forecast system (e.g., an ensemble Kalman filter or an ensemble of data assimilations). In our case, we did not have access to such high-quality ensemble forecasts; instead, we initialized the ensemble forecasts by drawing initial ensemble perturbations from a database of 4DVAR increments. We used 400 ensemble perturbations from the time period 21 November 2014 to 2 March 2015 (Fig. 2). For sensitivity tests with smaller ensemble sizes, we randomly selected a subset of increments over the same time period. The increments were added to the initial background state, and then 6-h forecasts were completed for each ensemble member.
During early stages of our experimentation, we also tried to use the flow-dependent ensemble that was cycled in NAVGEM using the ET method described in McLay et al. (2008) and Kuhl et al. (2013). However, we found that the performance of the LETLM with ET perturbations was significantly worse. The root-normalized-mean-square error [RNMSE in Eq. (7)] with the analysis increment ensemble was 0.313, compared to 0.505 with the ET ensemble—a 65% degradation. Our preliminary analysis suggested that this degradation was because the ET ensemble did not span the space of possible climatological increments within each influence volume. The results of the experiments with the ET-based ensemble are not shown in this paper. We concede that the optimal generation of the ensemble for the LETLM is still an open question, and we defer it to future investigations.
c. Error metrics





















d. Nonlinear index


e. Tuning methodology
To illustrate the LETLM approach to parameter optimization, we consider two “tuning” parameters that are available in our LETLM: 1) the horizontal length scale L of the localization window that determines the horizontal extent of the influence volume (red dots in Fig. 1) and 2) the pseudoinverse cutoff parameter
To find the optimal values of L and
Clearly, it would be highly impractical to conduct such extensive tuning every time one would want to use an LETM. Instead, we assume tuned profiles of Lopt(k) and
4. LETLM reference configuration
a. Reference configuration
To illustrate the performance of the LETLM, we assembled a reference configuration of the LETLM that provides a good balance between the accuracy and the computational efficiency. For the vertical influence volume, we chose zhalo = 4; for the ensemble size, we used the maximum number of analysis increments in our archive (400); the time step of the LETLM was fixed at 1 h, twice the time step of the nonlinear model; and the forecast time was 6 h. We chose the 1-h time step and 6-h period because the NAVDAS-AR 4DVAR needs to propagate observations backward in time in hourly batches over a 6-h DA window. We also felt that the 1-h time step allowed us to use relatively small influence volume (recall that the size of the influence volume grows with the length of the time step, which increases computational costs and complicates effective localization of the LETLM). The LETLM state variables included U, V, T, P, Z,

Optimally tuned profiles of (a),(c) horizontal halo region L and (b),(d) pseudoinverse cutoff β used in the reference configuration. Profiles are shown on pressure levels in (a),(b) and on the bottom 30 model levels (up to approximately 220 hPa) in (c),(d). Horizontal line in (a) indicates the location of the model level 30 (around 220 hPa). Vertical line in (b),(d) shows the default cutoff value of 1.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Optimally tuned profiles of (a),(c) horizontal halo region L and (b),(d) pseudoinverse cutoff β used in the reference configuration. Profiles are shown on pressure levels in (a),(b) and on the bottom 30 model levels (up to approximately 220 hPa) in (c),(d). Horizontal line in (a) indicates the location of the model level 30 (around 220 hPa). Vertical line in (b),(d) shows the default cutoff value of 1.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Optimally tuned profiles of (a),(c) horizontal halo region L and (b),(d) pseudoinverse cutoff β used in the reference configuration. Profiles are shown on pressure levels in (a),(b) and on the bottom 30 model levels (up to approximately 220 hPa) in (c),(d). Horizontal line in (a) indicates the location of the model level 30 (around 220 hPa). Vertical line in (b),(d) shows the default cutoff value of 1.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Parameters for reference LETLM configuration and ranges of tuning values.


b. Tuned profiles of optimal length and pseudoinverse cutoff
Figure 3 shows how the optimally tuned values of L and β varied vertically. L was, on average, around 750 km within the troposphere (Fig. 3c); however, it increased steadily in the stratosphere to 1500–1750 km above 1.0 hPa (Fig. 3a). The Courant–Friedrichs–Lewy (CFL) analysis of such large influence volumes suggests a maximum propagation speed of ~200 m s−1 in the troposphere and ~400 m s−1 in the lower mesosphere. We were surprised to obtain such large propagation speeds, which exceed the physical wind speeds in the atmosphere. In the stratosphere, they even exceed the speed of propagating gravity waves or the speed of sound (343 m s−1). In other words, in the stratosphere, the optimal size of the horizontal influence volume exceeded the length at which information can propagate in the volume. We speculate that such large empirically tuned horizontal halos resulted from the limited size of our vertical halo. The NAVGEM updates the entire vertical column at once. Hence, it is theoretically possible for variations at the bottom of the atmosphere to cause variations at the top of the atmosphere over a single time step. Our reference configuration for the LETLM can only accommodate information propagation over nine vertical levels. We speculate that the LETLM compensates for the lack of the vertical information by enlarging the horizontal halo region to include information correlated to the missing vertical information.
The term
c. Performance of the reference LETLM
We found that both the LETLM and the TLM show considerable skill, compared to persistence, in predicting evolution of the 4DVAR increment over the 6-h assimilation window. The growth of the vertically averaged errors (Fig. 4) shows the LETLM errors were consistently lower than the NOGAPS TLM errors over the entire forecast, except at t = 0 h. We note that while the initial TLM errors are exactly zero, the initial LETLM errors are small nonzero values. This is because the initial TLM field was initialized with the “true” perturbation, while the LETLM perturbation was projected into a subspace of a localized ensemble. Since the projection was not exact, there are slight differences relative to the truth at t = 0 h. Overall, the normalized errors favor the LETLM. At t = 6 h, the LETLM error was 0.303, while the NOGAPS TLM error was 0.340, an improvement of 11%.

Vertically integrated RNMSEs in propagation of the 4DVAR correction over 6 h [Eq. (7)]. Plotted are the magnitude of the propagated perturbation at forecast time tau (blue); errors under the assumption of a trivial TLM = I, equivalent to the 3DVAR case (red); errors for propagation of the perturbation using the LETLM (green); and errors for propagation of the perturbation using the NOGAPS TLM (black).
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Vertically integrated RNMSEs in propagation of the 4DVAR correction over 6 h [Eq. (7)]. Plotted are the magnitude of the propagated perturbation at forecast time tau (blue); errors under the assumption of a trivial TLM = I, equivalent to the 3DVAR case (red); errors for propagation of the perturbation using the LETLM (green); and errors for propagation of the perturbation using the NOGAPS TLM (black).
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Vertically integrated RNMSEs in propagation of the 4DVAR correction over 6 h [Eq. (7)]. Plotted are the magnitude of the propagated perturbation at forecast time tau (blue); errors under the assumption of a trivial TLM = I, equivalent to the 3DVAR case (red); errors for propagation of the perturbation using the LETLM (green); and errors for propagation of the perturbation using the NOGAPS TLM (black).
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
When we examine the globally averaged profiles of errors valid at t = 6 h (Figs. 5, 6), we find that the LETLM errors were roughly the same as the TLM errors in the boundary layer, but from ~970 to 100 hPa, the LETLM errors were smaller for all variables. (Note that the reference pressure labels on Figs. 5 and 6 were calculated using the NAVGEM 60-level hybrid sigma-pressure coefficients and assuming a surface pressure of 1000 hPa; all error calculations are actually performed on the model levels, not constant pressure levels.) From 100 to ~0.5 hPa, the TLM errors are generally smaller. When we compare in Fig. 7 the relative performance of the LETLM and TLM to the value of the NLI, we find that LETLM performed better in the areas (below 100 hPa) where the nonlinearity was high (NLI > 1.5), and the TLM performed better in the regions (above 100 hPa) where the model was more linear (NLI < 1.5). In the boundary layer where the model was dominated by nonlinearity (NLI > 4), both the TLM and the LETLM had similar poor performance. This behavior is consistent with the shallow water model results of Allen et al. (2017), which showed that the LETLM performed better relative to the TLM as the nonlinearity of the analysis increment increased [see Fig. 16b of Allen et al. (2017)].

Globally averaged profile of errors for (a) Eq. (6) and (b)–(d) Eq. (5) after 6 h.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Globally averaged profile of errors for (a) Eq. (6) and (b)–(d) Eq. (5) after 6 h.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Globally averaged profile of errors for (a) Eq. (6) and (b)–(d) Eq. (5) after 6 h.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

As in Fig. 5, but for the lowest 30 levels of the model.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

As in Fig. 5, but for the lowest 30 levels of the model.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
As in Fig. 5, but for the lowest 30 levels of the model.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

NLI [computed using Eq. (12)].
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

NLI [computed using Eq. (12)].
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
NLI [computed using Eq. (12)].
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
The correlation profiles

Average correlation of true and propagated perturbations after 6 h. (a) Performance from the top of the atmosphere to the surface with logarithm of pressure used as a vertical coordinate. (b) Troposphere (lower 30 model levels) with model levels used as vertical coordinates.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Average correlation of true and propagated perturbations after 6 h. (a) Performance from the top of the atmosphere to the surface with logarithm of pressure used as a vertical coordinate. (b) Troposphere (lower 30 model levels) with model levels used as vertical coordinates.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Average correlation of true and propagated perturbations after 6 h. (a) Performance from the top of the atmosphere to the surface with logarithm of pressure used as a vertical coordinate. (b) Troposphere (lower 30 model levels) with model levels used as vertical coordinates.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
We speculate that the comparatively poor performance of the NOGAPS TLM could be, in part, attributed to it being somewhat overdiffusive. For example, Fig. 9 shows that the globally averaged magnitudes of the perturbations in the NOGAPS TLM are lower than the magnitude of the true perturbation after 6 h. In contrast, the LETLM-propagated perturbations retain the correct magnitude almost everywhere in the profile, except in the boundary layer, where both the NOGAPS TLM and the LETLM significantly underpredict the magnitude of the perturbation. We further found support for our theory that NOGAPS TLM is too smooth, compared to the Truth or to the LETLM, when we examined the maps of the perturbations in the middle of the troposphere (~700 hPa) and stratosphere (~10 hPa). We found that the LETLM was able to capture the finescale structure in the U and V fields better than the NOGAPS TLM, in which the features appear to be somewhat smoothed out. For brevity, we present the extensive graphical comparisons of the maps for the Truth, the NOGAPS TLM, and the LETLM in the online supplemental material (Figs. S1 and S2 for 700 hPa and Figs. S3 and S4 for 10 hPa).

Average magnitude of the true perturbation (dotted line), the NOGAPS TLM (black line), and the LETLM (green line). (a)–(c) Performance from the top of the atmosphere to the surface with logarithm of pressure used as a vertical coordinate. (d)–(f) Troposphere (lower 30 model levels) with model levels used as vertical coordinates.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Average magnitude of the true perturbation (dotted line), the NOGAPS TLM (black line), and the LETLM (green line). (a)–(c) Performance from the top of the atmosphere to the surface with logarithm of pressure used as a vertical coordinate. (d)–(f) Troposphere (lower 30 model levels) with model levels used as vertical coordinates.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Average magnitude of the true perturbation (dotted line), the NOGAPS TLM (black line), and the LETLM (green line). (a)–(c) Performance from the top of the atmosphere to the surface with logarithm of pressure used as a vertical coordinate. (d)–(f) Troposphere (lower 30 model levels) with model levels used as vertical coordinates.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
One area where both TLMs had relatively poor performance was the prediction of temperature in the boundary layer. The maps of the perturbation at level 59 (close to the bottom of the boundary layer) show that the true perturbation develops small-scale features in a 6-h forecast, such as the large temperature perturbations over Africa (row two of Fig. 10). These features are absent in the original perturbation at 0 h (top row of Fig. 10), as well as in the 6-h forecast with either the NOGAPS TLM (third row from the top) or the LETLM (bottom row). In fact, the TLM- and LETLM-propagated fields look like smooth, evolved versions of the 0-h field. We speculate that these unresolved, intense features are likely due to the nonlinearity of the convection and the planetary boundary layer schemes in the NAVGEM. Recall that these high-intensity features are present in the original ensemble forecast, but our tuning experiments (Fig. 3) suggest that we need to increase the pseudoinverse cutoff parameter in the boundary layer in order to filter out some features of the ensemble forecast that cannot be predicted by the linear forecast operator.

Horizontal maps of the U, V, and T perturbations plotted at level 59 (almost at the surface) at (top) the initial time and (bottom three rows) 6 h (Truth is in row 2, TLM is in row 3, and LETLM is in row 4). Perturbations are shown for the reference configuration of the system.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Horizontal maps of the U, V, and T perturbations plotted at level 59 (almost at the surface) at (top) the initial time and (bottom three rows) 6 h (Truth is in row 2, TLM is in row 3, and LETLM is in row 4). Perturbations are shown for the reference configuration of the system.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Horizontal maps of the U, V, and T perturbations plotted at level 59 (almost at the surface) at (top) the initial time and (bottom three rows) 6 h (Truth is in row 2, TLM is in row 3, and LETLM is in row 4). Perturbations are shown for the reference configuration of the system.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Finally, we examined the temporal evolution of the U perturbation along a vertical cross section taken along the 267.5° longitude (Fig. 11). The cross section shows a variety of scales captured by both the TLM and the LETLM, including more rapid evolution of the field at the upper levels. This may be related to resolved unbalanced modes (gravity waves) that are present in the system. These fast-moving features require larger length scales, as seen in tuning of the reference configuration. The LETLM and the TLM both capture the temporal variability in the cross sections well.

Hourly cross sections of the (left) Truth, (middle) TLM-propagated, and (right) LETLM-propagated perturbations. Cross sections of perturbations are shown for the reference configuration of the system and are taken along the 267.5°E longitude.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Hourly cross sections of the (left) Truth, (middle) TLM-propagated, and (right) LETLM-propagated perturbations. Cross sections of perturbations are shown for the reference configuration of the system and are taken along the 267.5°E longitude.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Hourly cross sections of the (left) Truth, (middle) TLM-propagated, and (right) LETLM-propagated perturbations. Cross sections of perturbations are shown for the reference configuration of the system and are taken along the 267.5°E longitude.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Overall, we conclude that the reference configuration of the LETLM with offline tuning and zhalo = 4 competes well with the TLM in terms of the global error profiles, horizontal maps, and cross sections. In the next section, we further examine the sensitivity of the LETLM to various parameters in order to provide more insight into LETLM behavior and further guidance in tuning the LETLM for particular user needs.
5. Sensitivity tests for the LETLM
a. Sensitivity to horizontal length
In the reference configuration examined in section 4, we simultaneously tuned L and

Sensitivity to LETLM parameters with respect to the reference configuration [Eq. (11)]; positive values of error show degradation, and negative values show improvement over the reference configuration. Plotted are globally averaged vertical RNMSE profiles for (a) localization scale L, (b) vertical localization scale zhalo, (c) pseudocutoff parameter β, and (d) the ensemble size nens.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Sensitivity to LETLM parameters with respect to the reference configuration [Eq. (11)]; positive values of error show degradation, and negative values show improvement over the reference configuration. Plotted are globally averaged vertical RNMSE profiles for (a) localization scale L, (b) vertical localization scale zhalo, (c) pseudocutoff parameter β, and (d) the ensemble size nens.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Sensitivity to LETLM parameters with respect to the reference configuration [Eq. (11)]; positive values of error show degradation, and negative values show improvement over the reference configuration. Plotted are globally averaged vertical RNMSE profiles for (a) localization scale L, (b) vertical localization scale zhalo, (c) pseudocutoff parameter β, and (d) the ensemble size nens.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

As in Fig. 12, but for the bottom 30 model layers.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

As in Fig. 12, but for the bottom 30 model layers.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
As in Fig. 12, but for the bottom 30 model layers.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Sensitivity of the globally averaged RNMSE to tuning parameters (a) localization scale L, (b) vertical localization scale zhalo, (c) pseudocutoff parameter β, and (d) the ensemble size nens. Horizontal dotted red line shows the RNMSE score for the NOGAPS TLM, and horizontal dotted black line shows the RNSME score of the reference configuration. The optimal value is indicated by a vertical tick mark.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Sensitivity of the globally averaged RNMSE to tuning parameters (a) localization scale L, (b) vertical localization scale zhalo, (c) pseudocutoff parameter β, and (d) the ensemble size nens. Horizontal dotted red line shows the RNMSE score for the NOGAPS TLM, and horizontal dotted black line shows the RNSME score of the reference configuration. The optimal value is indicated by a vertical tick mark.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Sensitivity of the globally averaged RNMSE to tuning parameters (a) localization scale L, (b) vertical localization scale zhalo, (c) pseudocutoff parameter β, and (d) the ensemble size nens. Horizontal dotted red line shows the RNMSE score for the NOGAPS TLM, and horizontal dotted black line shows the RNSME score of the reference configuration. The optimal value is indicated by a vertical tick mark.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
b. Sensitivity to vertical halo
To examine the sensitivity to vertical localization, we varied the number of adjacent levels (above and below) that are allowed to influence the central level. We ran with zhalo values of 0, 2, 4, 6, and 8, corresponding to a vertical localization of no larger than 1, 5, 9, 13, and 17 levels. In each case, we fixed
c. Sensitivity to pseudoinverse cutoff
We next examine the sensitivity to the pseudoinverse cutoff
d. Sensitivity to ensemble size
The LETLM was run with ensemble sizes of 40–400 in steps of 40 ensemble members. Figures 12d, 13d, and 14d show that the skill of the LETLM steadily increases as more ensemble members are added. It also appears that larger ensembles will yield even higher skill, as the RMS errors have not saturated yet at 400 ensemble members—the maximum number of ensembles that we tried (see Fig. 14d). We also found that addition of ensemble members was more effective at reducing errors above the troposphere (cf. the slopes of green and blue lines that show the RMSE below and above 100 hPa). The results in Fig. 14d suggest that even higher LETLM performance can be achieved either by using a larger ensemble (which is expensive) or by possibly using a higher-quality ensemble that can capture more information using fewer ensemble members.
In our prior work, Frolov and Bishop (2016) showed that the fidelity of the LETLM is improving steadily as the ensemble size is increased. Bishop et al. (2017) provided a mathematical proof that the LETLM is guaranteed to be precisely accurate when the rank of the ensemble within the influence volume is equal to the number of variables within the influence volume and when the ensemble perturbations are small enough to be governed by linear dynamics. As a crude measure of the degrees of freedom in the stencil, we compared the number of grid points in the influence volume with the number of ensemble members (Fig. 15). We find that the two numbers were comparable in the troposphere (with the exception of the polar regions, where the influence volume was as large as 2000 grid points). However, in the upper stratosphere and mesosphere, the number of points in the influence volume was significantly larger than the number of ensemble members (~800–2000 outside the polar regions).

Size of the influence volume (number of horizontal and vertical grid points) computed for the average tuning profile used in the verification tests in section 6.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Size of the influence volume (number of horizontal and vertical grid points) computed for the average tuning profile used in the verification tests in section 6.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Size of the influence volume (number of horizontal and vertical grid points) computed for the average tuning profile used in the verification tests in section 6.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
e. Sensitivity to the choice of state variables
The LETLM can be used to propagate any combination of state variables available in the nonlinear NAVGEM. We tested the LETLM with zhalo = 4 and nens = 400 for a number of different configurations of state variables (U, V, T, P, Z,

Sensitivity to the choice of the model variables. Sensitivities are computed with respect to the reference configuration using Eq. (11). The dotted black line indicates the NOGAPS TLM error relative to the reference configuration.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Sensitivity to the choice of the model variables. Sensitivities are computed with respect to the reference configuration using Eq. (11). The dotted black line indicates the NOGAPS TLM error relative to the reference configuration.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Sensitivity to the choice of the model variables. Sensitivities are computed with respect to the reference configuration using Eq. (11). The dotted black line indicates the NOGAPS TLM error relative to the reference configuration.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
f. Sensitivity to physics in the LETLM
One of the main difficulties of developing traditional TLMs is that the physics package of the atmospheric model (responsible for processes like mixing, radiative transfer, and parameterization of moist processes) is usually rapidly evolving, highly nonlinear, and often lacks a formal derivative because of conditional statements in the code that cannot be differentiated. In contrast, the LETLM has direct access to the entire physics package of the nonlinear model. Therefore, any changes to the nonlinear model will automatically be included in the LETLM. As with conventional TLMs, however, it is necessary to test the impact of each change to the nonlinear model on the performance of the TLM (Janisková and Lopez 2012). While a complete examination of the sensitivity of the LETLM to various physics packages is beyond the scope of the current study, we briefly show the impact of turning the physics and radiation packages off (i.e., running a dry adiabatic core).






Sensitivity of (a),(e) U; (b),(f) V; (c),(g) T; and (d),(h) specific humidity to the inclusion of physics and radiation in the ensemble calculations, calculated using Eq. (14).
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Sensitivity of (a),(e) U; (b),(f) V; (c),(g) T; and (d),(h) specific humidity to the inclusion of physics and radiation in the ensemble calculations, calculated using Eq. (14).
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Sensitivity of (a),(e) U; (b),(f) V; (c),(g) T; and (d),(h) specific humidity to the inclusion of physics and radiation in the ensemble calculations, calculated using Eq. (14).
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Figure 17 shows that surface winds were improved by as much as 50% by the inclusion of the physics, while the tropospheric improvement in winds was around 10%–20%. A minimum in error reduction was in the lower stratosphere, while larger improvements occur near the stratopause (1.0 hPa). The
6. Performance of the tuned LETLM for other increments
While the LETLM with full tuning has been shown to provide skill throughout the atmosphere, in an operational setting, it would be helpful to have pretuned profiles of L and β. The exact method of tuning and frequency of updating these profiles will need to be tested in detail. Here, we test a simple method in which full tuning is performed on two different dates at the beginning (0600 UTC 20 November 2014) and end (0000 UTC 4 March 2015) of the date range used for the ensemble initialization (Fig. 2). The optimal L and β profiles are slightly different for the two cases, but show broadly similar results. We then averaged the L and β profiles from these two offline tuning tests and used these to run online tests for seven other times picked randomly from 5 to 15 March 2015.
The resulting error profiles [relative to the corresponding TLM forecast errors, using Eq. (11)] are provided in Fig. 18, with colors corresponding to the times (in YYYYMMDD format) indicated in the top panel. The LETLM errors are consistently lower than the TLM errors over approximately the same vertical range as the reference profile (Figs. 5, 6). We computed the statistical significance of the mean profiles using the Student’s t test and found that the LETLM was significantly better (95% confidence level) at the surface, over a broad layer from model level 55 (~970 hPa) to level 30 (~200 hPa). The TLM errors were smaller than the LETLM errors from ~70 hPa to the model top. While the LETLM does not beat the TLM at these levels, the skill of the LETLM relative to persistence for these tests is significant and comparable to that of the reference configuration in Fig. 5a. Further tuning tests are required to understand how the LETLM responds in different seasons, but these preliminary results show promise that the LETLM will perform well, even with coarse tuning.

Verification tests with seven different analysis increments. (a) RNMSE results relative to the error in NOGAPS TLM [Eq. (11); negative values indicate improvement compared to the NOGAPS TLM and positive values, deterioration]. Black dots indicate levels where LETLM was significantly better (with 95% confidence using t test) than the NOGAPS TLM. Red dots indicate levels where the NOGAPS TLM was significantly better (with 95% confidence). (b) As in (a), but highlights the errors in the troposphere by using model levels as vertical coordinates. In (a) dates for each increment are shown in color labels.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1

Verification tests with seven different analysis increments. (a) RNMSE results relative to the error in NOGAPS TLM [Eq. (11); negative values indicate improvement compared to the NOGAPS TLM and positive values, deterioration]. Black dots indicate levels where LETLM was significantly better (with 95% confidence using t test) than the NOGAPS TLM. Red dots indicate levels where the NOGAPS TLM was significantly better (with 95% confidence). (b) As in (a), but highlights the errors in the troposphere by using model levels as vertical coordinates. In (a) dates for each increment are shown in color labels.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
Verification tests with seven different analysis increments. (a) RNMSE results relative to the error in NOGAPS TLM [Eq. (11); negative values indicate improvement compared to the NOGAPS TLM and positive values, deterioration]. Black dots indicate levels where LETLM was significantly better (with 95% confidence using t test) than the NOGAPS TLM. Red dots indicate levels where the NOGAPS TLM was significantly better (with 95% confidence). (b) As in (a), but highlights the errors in the troposphere by using model levels as vertical coordinates. In (a) dates for each increment are shown in color labels.
Citation: Monthly Weather Review 146, 7; 10.1175/MWR-D-17-0315.1
7. Computational efficiency
a. Profiling of the LETLM computations
Our profiling of the LETLM computations (Table 3) shows that computing the pseudoinverse of the ensemble covariance matrix
Profiling of the LETLM computation. Note: The computational times were obtained for the reference configuration of the LETLM.


b. Parallel scalability
We conducted two types of scalability tests. In the weak scalability test, we fixed the computation hardware but increased the number of computations in the LETLM by varying LETLM tuning parameters. In the strong scalability test, we fixed the size of the problem and used progressively larger numbers of computational processing units.




Weak scalability results.


The results of the strong scalability tests showed that the LETLM was capable of close-to-perfect scalability up to 700 processing cores. The scalability at high CPU counts was limited by the startup costs of the OpenMP threads (as was evident by better scalability when more MPI tasks were used instead of OpenMP threads). It is possible to further optimize these startup costs. We also suggest that the strong scalability will improve even further once we use a higher-resolution grid because there will be more work to be done, and the startup costs would be less important.
We also found that the LETLM strong scalability was significantly better than that of the original model (the NAVGEM only scaled to 10 cores). We explain such poor scalability of the NAVGEM by the very small size of our global grid. From our experience, we know that it is possible to scale NAVGEM further on high-resolution model simulations. We explain the positive results with LETLM strong scalability by the fact the LETLM almost completely avoids global communications.
Overall, we found that the cost of the LETLM (in terms of CPU clock) was dominated by the cost of generating ensemble forecasts (two orders of magnitude larger than the cost of a single integration of the stand-alone NOGAPS TLM or a stand-alone NAVGEM). The cost of the LETLM computation was comparable to or smaller than the cost of the ensemble forecast and depended greatly on the configuration of the LETLM [see weak scalability results in Table 4 and Eq. (15)]. We believe that it should be possible to further reduce the cost of the LETLM computation by a factor of 5–10. We discuss these strategies in the following section.
8. Summary and conclusions
This paper presents a first demonstration of the ensemble-based TLM performance in a realistic, albeit low-resolution, model of the global atmosphere. Our results show that the LETLM outperformed the NOGAPS TLM on average (by 10% over a 6-h forecast). The NOGAPS TLM is the TLM currently used by the hybrid-4DVAR system run in operations by the U.S. Navy’s Fleet Numerical Meteorology and Oceanography Center. The superior performance was most prominent for winds in the troposphere (about 20% improvement). Both TLMs had similarly poor skill for forecasting temperature perturbations in the planetary boundary layer. The NOGAPS TLM had superior skill in the stratosphere and mesosphere. Comparisons of these skill scores with the measure of model nonlinearity are not inconsistent with the speculation that the NOGAPS TLM was better when the model was only slightly nonlinear [e.g., in the stratosphere (NLI < 1.5)]. The LETLM was better in more nonlinear regimes (NLI between 2 and 4), and both linear models had poor skill in predicting strongly nonlinear growth of small perturbations in the planetary boundary layer (NLI of 4–8).
While the superior skill of the LETLM in the troposphere is encouraging, it should be evaluated in a proper context. Our skill comparisons were against a NOGAPS TLM—a TLM that was developed for a predecessor of the current NAVGEM and has not seen any significant updates in over a decade. More recently, an effort has been undertaken to develop a more up-to-date version of the TLM for the NAVGEM. This newer version shows similar skill improvement to the LETLM when compared against the NOGAPS TLM (about 15%–20% in the troposphere). Furthermore, an experience with a state-of-the-art TLM at ECMWF (P. Lopez 2017, personal communication) suggests that a continued effort at improving a traditional TLM leads to significant skill gains by the TLM, even in areas where both of our TLMs failed, such as the planetary boundary layer. We expect that similar gains can be achieved if a continuous effort is dedicated to the improvements in the ensemble system that drives the LETLM. Taking this larger context into account, it is fair to say that the LETLM is a promising technology that can deliver skill comparable to the skill of state-of-the-art TLMs.
Our sensitivity studies showed that the LETLM was most sensitive to the construction and the size of the ensemble. The ability of the LETLM to propagate analysis corrections was higher when the initial perturbations were obtained from an archive of 4DVAR analysis corrections than when using the initial perturbations that are typically used for ensemble forecasting (the ET initial perturbations). The size of the ensemble also mattered. The performance of the system continued improving up to the maximum size of the ensemble attempted (400 members). It is very likely that larger ensembles will result in even higher skill. We also speculate that it should be possible to construct more effective ensembles that can yield similar skill with fewer members.
Our sensitivity results also showed that allowing for vertical tuning of the horizontal halo L and pseudoinverse cutoff β resulted in moderate improvement of the LETLM performance (5.6% and 2.6%, respectively). The impact of tuning the vertical halo (zhalo) was most prominent in the stratosphere and the boundary layer. We found that tuning of zhalo interacted with the optimal length of the horizontal halo (bigger vertical halos required smaller horizontal halos). In general, the horizontal halo increased with height (from 750 km in the troposphere to 1750 km on top of the stratosphere). We speculate that the larger horizontal halo is better able to represent the large-scale rapidly propagating waves in the upper stratosphere. Recent work (to be presented in future publications) has shown that a further reduction in the LETLM errors is possible by using a nonuniform influence volume, where the current cylindrical influence volume (of height zhalo and radius L) is complemented by a pencil-like column that extends the influence volume directly above and below the central patch point. This new influence volume shape (we call it a “spinning top”) is especially effective at reducing errors at the model top and in the boundary layer, where information can propagate rapidly in the vertical due to convection or radiation processes.
Our results also showed that including physics in the LETLM computation yielded positive results through most of the vertical column (improvement of 23%, compared with the LETLM based on the ensemble without parameterizations of subgrid-scale processes and radiation). These sensitivities to physics are in qualitative agreement with previous findings from the physics-enabled TLMs developed at ECMWF (Janisková and Lopez 2012). However, we suspect that the impact of physics will increase with higher model resolution. The experience at ECMWF also suggests that to achieve strong skill scores in the boundary layer, it might be necessary to develop versions of the model ensemble that have simplified physics in the boundary layer that will improve linear predictability of the system at short time scales.
The profiling of the computational costs in the LETLM showed that the overall cost was dominated by the ensemble forecast, with the cost of the LETLM computation being similar or lower than the cost of the ensemble integration. However, all operational NWP centers are already committing large resources for ensemble forecast integration. For example, the Canadian Meteorological Service routinely performs forecasts with as many as 500 ensemble members at 50-km resolution. The LETLM can leverage these existing computations without adding significantly to the cost of the overall system. In terms of the LETLM computation, the majority of the cost was associated with the matrix inverse (the size of the ensemble) at each of the model grid points. There are several strategies that can allow us to further reduce the cost of the LETLM computation, which we plan to examine in the follow-up publications. These strategies include computing the matrix inverse less frequently (e.g., at every other grid point), increasing the size of the vertical or horizontal computational patch (currently, every grid point forms a unique patch), and developing methods to construct a smaller subset of ensemble members that can be as effective as our current climatological ensemble.
However, regardless of the success of the abovementioned optimization, the LETLM computation has a unique property that will allow it to significantly reduce the cost of computations where the TLM and its adjoint are applied iteratively (such as in 4DVAR). In fact, the LETLM matrix can be precomputed before the 4DVAR executes and stored as a sparse matrix that can then be invoked by the 4DVAR solver iteratively at almost zero cost. Furthermore, the sparse matrix multiplication is supported by robust libraries and is likely to scale better than a traditional TLM computation. This possibility to drastically reduce the cost of the 4DVAR computation (or enable the efficient implementation of the perturbed observation 4DVAR system) is one of the most appealing properties of the proposed system. It is also likely that the speed-up based on the precomputed LETLM can be realized even in 4DVAR systems that are configured to run using an outer-loop configuration. In such systems, it might be beneficial to recompute the LETLM between the outer loops using a new set of ensemble forecasts that are conditioned based on the updated best guess from the previous outer loop. This speed-up may be possible because it is usually more efficient to compute ensemble forecasts in parallel rather than computing multiple sequential integrations of the TLM and adjoint models.
9. Future work
The feasibility of an NWP system based on the LETLM will depend on the ability of the LETLM to perform well in high-resolution test cases. We would like to provide a few remarks to the readers on how we think the findings of this study will generalize to high-resolution cases. As the resolution of the model increases, the number of grid points within an influence volume of a given size will scale by approximately
Another important aspect for the success of the LETLM in future operational-scale NWP is the ability to construct skillful ensembles that can support accurate LETLM computations. In this paper, we used a very crude ensemble that was initialized using an archive of historic analysis increments. We suspect that better and more effective ensembles can be constructed. Our results also shed light on some of the concerns raised by anonymous reviewers. One of the concerns was that an ensemble of forecasts needs to span the space of an arbitrary analysis correction. Clearly, it is impossible for any limited-size ensemble to exactly include an arbitrary analysis correction in a high-dimensional NWP model. However, our experience with the crude ensemble generation scheme used here and the EnKF-based scheme used in Allen et al. (2017) suggests that the LETLM localization method is effective at extracting the local dynamics of the system. We suspect that to construct an effective LETLM, it is sufficient if the ensemble perturbations are effective at exciting dominant dynamical modes for each of the localization volumes.
Finally, more work might be needed to understand the ability of LETLM to deal with unbalanced modes in either the ensemble forecast or in the increment that it propagates. In our earlier work (Allen et al. 2017), we showed that the normal mode initialization of the ensemble forecasts improves the skill of the LETLM, possibly by filtering out spurious gravity waves in the perturbation forecast that are generated due to the presence of unbalanced modes in the perturbation. We also showed that increasing the value of the regularization parameter β can act in a similar way by removing smaller-scale wavenumbers from the LETLM. In this paper, we had much less control on our ability to filter out unbalanced modes from either the forecast or the initial analysis perturbation. In fact, the only control we had was through a crude geostrophic balance in the 4DVAR error covariance. In other words, it is safe to say that both the analysis increment and the initial conditions for the ensemble forecasts had unbalanced modes. However, our results show that the LETLM was effective even in the presence of unbalanced modes (e.g., see rapid propagation of an unbalanced mode in Fig. 11 above 10 hPa and centered on 30°N). We did notice that we had to increase the size of the influence volume considerably in the stratosphere to accommodate propagation of the rapidly evolving gravity waves.
Acknowledgments
This work was funded by the U.S. Office of Naval Research. Sergey Frolov and Craig H. Bishop acknowledge support from the Naval Research Base Program PE-0601153N. Douglas R. Allen and Karl W. Hoppel acknowledge support from Office of Naval Research base funding via Task BE-033-02-42. David D. Kuhl and Karl W. Hoppel acknowledge support from Office of Naval Research base funding via Task BE-435-050. Authors thank Philippe Lopez of ECMWF for useful feedback. All are grateful for the access to the Department of Defense high-performance computing resources that enabled us to conduct this research. We also thank three anonymous reviewers for the most thorough review of our manuscript.
APPENDIX
Efficient Implementation of the Matrix Inverse for LETLM











































The Cholesky factorization method is computationally more efficient (by a factor of 2 in our experiments) than the singular value method because it is less expensive to compute the square root of the matrix than the singular spectrum and vectors of
The template for the LETLM computations is available in supplementary material to this article.
REFERENCES
Allen, D. R., C. H. Bishop, S. Frolov, K. W. Hoppel, D. D. Kuhl, and G. E. Nedoluha, 2017: Hybrid 4DVAR with a local ensemble tangent linear model: Application to the shallow-water model. Mon. Wea. Rev., 145, 97–116, https://doi.org/10.1175/MWR-D-16-0184.1.
Bishop, C. H., S. Frolov, D. R. Allen, D. D. Kuhl, and K. Hoppel, 2017: The local ensemble tangent linear model: An enabler for coupled model 4DVAR. Quart. J. Roy. Meteor. Soc., 143, 1009–1020, https://doi.org/10.1002/qj.2986.
Bowler, N. E., and Coauthors, 2017: The effect of improved ensemble covariances on hybrid variational data assimilation. Quart. J. Roy. Meteor. Soc., 143, 785–797, https://doi.org/10.1002/qj.2964.
Buehner, M., P. L. Houtekamer, C. Charette, H. L. Mitchell, and B. He, 2010: Intercomparison of variational data assimilation and the ensemble Kalman filter for global deterministic NWP. Part II: One-month experiments with real observations. Mon. Wea. Rev., 138, 1567–1586, https://doi.org/10.1175/2009MWR3158.1.
Eckermann, S. D., and Coauthors, 2009: High-altitude data assimilation system experiments for the northern summer mesosphere season of 2007. J. Atmos. Sol.-Terr. Phys., 71, 531–551, https://doi.org/10.1016/j.jastp.2008.09.036.
Frolov, S., and C. H. Bishop, 2016: Localized ensemble-based tangent linear models and their use in propagating hybrid error covariance models. Mon. Wea. Rev., 144, 1383–1405, https://doi.org/10.1175/MWR-D-15-0130.1.
Hogan, T. F., and Coauthors, 2014: The Navy Global Environmental Model. Oceanography, 27, 116–125, https://doi.org/10.5670/oceanog.2014.73.
Janisková, M., and P. Lopez, 2012: Linearized physics for data assimilation at ECMWF. ECMWF Tech. Memo. 666, 26 pp., https://www.ecmwf.int/sites/default/files/elibrary/2012/10163-linearized-physics-data-assimilation-ecmwf.pdf.
Kleist, D. T., and K. Ide, 2015: An OSSE-based evaluation of hybrid variational–ensemble data assimilation for the NCEP GFS. Part II: 4DEnVar and hybrid variants. Mon. Wea. Rev., 143, 452–470, https://doi.org/10.1175/MWR-D-13-00350.1.
Kuhl, D. D., T. E. Rosmond, C. H. Bishop, J. McLay, and N. L. Baker, 2013: Comparison of hybrid ensemble/4DVar and 4DVar within the NAVDAS-AR data assimilation framework. Mon. Wea. Rev., 141, 2740–2758, https://doi.org/10.1175/MWR-D-12-00182.1.
Lorenc, A. C., N. E. Bowler, A. M. Clayton, S. R. Pring, and D. Fairbairn, 2015: Comparison of hybrid-4DEnVar and hybrid-4DVar data assimilation methods for global NWP. Mon. Wea. Rev., 143, 212–229, https://doi.org/10.1175/MWR-D-14-00195.1.
McLay, J. G., C. H. Bishop, and C. A. Reynolds, 2008: Evaluation of the ensemble transform analysis perturbation scheme at NRL. Mon. Wea. Rev., 136, 1093–1108, https://doi.org/10.1175/2007MWR2010.1.
Rosmond, T. E., 1997: A technical description of the NRL adjoint modeling system. NRL Publ. NRL/MR/7532-97-7230, 57 pp., http://www.dtic.mil/dtic/tr/fulltext/u2/a330960.pdf.
Rosmond, T. E., and L. Xu, 2006: Development of NAVDAS-AR: Non-linear formulation and outer loop tests. Tellus, 58A, 45–58, https://doi.org/10.1111/j.1600-0870.2006.00148.x.
Tikhonov, A., and V. Arsenin, 1977: Solutions of Ill-Posed Problems. Winston and Sons, 258 pp.
Xu, L., T. Rosmond, and R. Daley, 2005: Development of NAVDAS-AR: Formulation and initial tests of the linear problem. Tellus, 57A, 546–559, https://doi.org/10.3402/tellusa.v57i4.14710.