P3.4 THE LOCAL ENSEMBLE OF THE UNIVERSITY OF MARYLAND

Istvan Szunyogh ∗ Eric J. Kostelich,† Gyorgyi Gyarmati, Brian R. Hunt, Eugenia Kalnay, Edward Ott, Dhanurjay Patil, and James A. Yorke University of Maryland, College Park, MD 20742

1. Introduction results are shown in Szunyogh et al. (2004), the present paper summarizes the final results The time has come when ensemble-based Kalman of this recently completed task. filter schemes can be considered for implementation on operational weather forecast • Implementation of the LEKF on the Regional systems in the foreseeable future. For the first Spectral Model (RSM) of NCEP. The RSM has time, an ensemble Kalman filter has been reported been selected since it has the most consis- to break even with a sophisticated operational 3D- tent dynamics with the GFS among all regional Var system (Houtekamer et al 2004), to outper- models. In our formulation, the regional analy- form the NCEP 3D-Var in reconstructing the state sis is essentially a refinement of the global anal- of the mid-troposphere from surface pressure ob- ysis prepared for the GFS. This implementa- servations (Whitaker et al. 2004), and to be ef- tion of the LEKF is currently under testing, and ficient in assimilating simulated and real Doppler- we hope to report on some preliminary results radar observations of convective systems (Snyder soon. and Zhang 2003; Zhang et al 2004; Dowell et al 2004). • Development and testing of methods to com- The present paper reports on the current status pensate for model errors via modification of the of the 4-dimensional Local Ensemble Kalman Filter data assimilation technique. data assimilation system (4D LEKF) developed by • Assimilating real observations (including a our interdisciplinary team at the University of Mary- large number of remotely sensed wind obser- land. Our plan has been to develop a largely model vations) with the 4D LEKF into both the GFS independent analysis system through completation and RSM forecast systems . This step is in of the following tasks: progress, and we hope to report on some re- • Derivation of the basic LEKF algorithm, and sults in the coming year. extensive testing on the Lorenz-96 model (us- ing from 40 to 120 dynamical variables). This The LEKF scheme is an ensemble square-root step has been completed and the results are filter (Tippett et al. 2002): one first obtains an es- reported in Ott et al. (2002 and 2004). timate of the most likely state of the atmosphere and an analysis error that de- • Extending the theory, and the low-order model scribes the uncertainty in the best estimate. Then experiments, to the case when asynchronous an ensemble of analyses is generated that is cen- observations are assimilated at a fixed assim- tered on the most likely state and is representa- ilation time. The resulting scheme, called 4D tive of the uncertainty reflected by the analysis er- LEKF, and the results for low-order model ex- ror covariance matrix. A distinguishing feature of periments are described in Sauer et al. (2004) the LEKF is that it solves the Kalman filter equa- and Hunt et al. (2004). tions locally in model grid space; other square- • Testing the largely model independent com- root filters solve the Kalman filter equation locally puter code of the LEKF on an operational global in observation space (Bishop et al. 2001; Ander- numerical weather forecast model. For this son 2001, Whitaker and Hamill 2002). More pre- purpose we selected the Global Forecast Sys- cisely, the LEKF obtains the analysis at the differ- tem (GFS) of the National Centers for Environ- ent grid points independently, using all observations mental Prediction (NCEP). While preliminary that are thought to improve the analysis at the indi- vidual grid points. In this scheme, the same obser- ∗ Corresponding author address: University of Maryland, vation may be used to obtain the analysis at multiple IPST, CSS Building, College Park, MD 20742-2431, USA †Current address: Arizona State University, Tempe, AZ grid points. On the other hand, sequential schemes 85287, USA assimilate the observations one by one (or by small groups when the errors between the observations [A simple technique to extend the scheme to the are correlated), iteratively updating the state esti- assimilation of asynchronous observations is pre- mate at those grid-points, where the accuracy of the sented in (Hunt et al. 2004).] analysis is thought be positively affected by a given observation (or group of observations). 2.1 Global and local background vectors We speculate that solving the Kalman filter A k + 1-member ensemble (k ≥ 1) of global back- equations locally in grid space may be computation- b(i) ground state vectors, xg , i = 1, 2, ··· , k + 1, is ally advantageous and, at the same time, may not obtained by integrating the forecast model started noticably degrade the accuracy of the assimilation. from a k + 1-member ensemble of analysis fields The local analyses can be processed in parallel, in- created in the previous analysis cycle. volve relatively small matrices, and treat all data si- For each grid point p, we define a correspond- multaneously. These features suggest the potential ing local region that consists of all grid points for the LEKF method in rapidly and efficiently assim- within a suitably prescribed neighborhood of p. Let ilating large amounts of data (e.g., as will become xl (m, n, o) be the d-dimensional local vector repre- available from future satellite observing systems.) senting the model state within the local region cen- In what follows, we first provide a short sum- tered at the grid point (m, n, o). The construction of mary of the LEKF algorithm (Section 2), then ex- this local vector is a linear mapping L of the vec- plain the implementation of the scheme on the tor xg that represents the state of the model in the NCEP GFS (Section 3). This implementation is space defined by the global three-dimensional grid tested under the perfect model hypothesis, i.e., by space. Since all the analysis operations take place assuming that a model run provides a perfect rep- at a fixed time t and are repeated for all local re- resentation of the true evolution of the atmosphere, gions, henceforth we suppress the dependence of making possible the generation of simulated obser- all vectors and matrices on t, m, n, and o. vations (with known error statistics) and the exact The local background error covariance matrix computation of analysis and forecast errors (Sec- b Pl and the most probable local background state tion 4). Experiments are carried out for different b x¯ l are derived from the k + 1-member ensemble of size ensembles and for varying observational data b(i) global state field vectors xg , i = 1, 2, ··· , k +1. The coverage (Section 5). The accuracy of the analy- most probable local state is estimated by sis scheme is measured by the root--square distance between the true states and the analy- " k+1 # b −1 b(i) ses. The computational efficiency is measured by x¯ l = L (k + 1) ∑ xg , (1) the wall-clock time needed to complete the anal- i=1 ysis. The results of this experiment indicate that while the d × d local background error covariance the LEKF may become an operationally feasible b matrix Pl is estimated by scheme (Section 6). k+1 T b −1 b(i)  b(i) Pl = k ∑ δxl δxl , (2) 2. Local ensemble Kalman filter i=1

A detailed description and mathematical justifi- where the superscript T denotes transpose and cation of the different components of the LEKF δxb(i) = Lxb(i) − x¯ b. (3) scheme can be found in Ott et al. (2002 and 2004). l g l Here we provide only a brief algorithmic summary b We can express Pl in terms of the d ×(k +1) matrix, needed to understand the implementation of the b −1/ 2 h b(1) b(2) b(k+1)i scheme on the NCEP GFS. The version of the Xl = (k) δxl | δxl | · · · | δxl , (4) scheme that we describe assumes that the rank of the background and analysis covariance matrices as b b bT is k when the ensemble has k + 1 members. Ott Pl = Xl Xl . (5) et al. (2004) describes a more general formulation 2.2 Projection onto the k-dimensional analysis that allows for a reduction of the rank. We note, space that while we use one particular form of the matrix square root, other square roots can also be consid- By using a k +1-member ensemble, we assume that ered (Tippett et al 2002; Ott et al. 2004). Finally, we an estimate of the background covariance matrix of consider the case where all observations collected rank k is sufficient to obtain accurate analyses. Ex- for the current analysis are taken at the same time. perience accumulated by others (Houtekamer and a Mitchell 2000; Keppenne and Rienecker 2002) sug- The most probable value of ∆xˆ l is gest that k + 1 may be as small as 40. For the pur-   ˆ a ˆ a ˆ T −1 o b pose of subsequent computations, we consider the ∆x¯ l = Pl Hl R yl − H x¯ l . (13) coordinate system of the k-dimensional space de- ˆ termined by the k orthonormal eigenvectors {u(j)} Here Hl = Hl Q is the Jacobian matrix of partial of Pb, which we use to form the internal coordinate derivatives of the observation operator Hl (evalu- l ated atx ¯ b); thus, Hˆ = H Q maps variables from system for the k-dimensional local analysis space. l l l b the the k-dimensional representation of the analysis Since Pl has rank k, it has k positive eigenvalues a T a analysis to the observation space. ∆xˆ l = Q ∆xl λ(1) ≥ λ(2) ≥ ... ≥ λ(r) ≥ · · · ≥ λ(k) > 0. (6) is the analysis increment in the k-dimensional anal- ˆ a ysis space, and Pl is the analysis error covariance Thus, matrix in the same k-dimensional space.(If there k are s scalar observations in the local region at the b (j) (j) (j) T Pl = ∑ λ u (u ) . (7) o analysis time,y ¯ l is s-dimensional and the rectangu- j=1 ˆ a lar matrix Hl is s × d). In Eq. (13), Pl is determined Since the size of the ensemble (k + 1) is envisioned from the usual Kalman filter equations (e.g., Kalnay b to be much less than the dimension d of xl , the 2003), but restricted to the k-dimensional internal (j) computation of the basis vectors {u } is most ef- coordinate system, we have ficiently done in the basis of the ensemble vectors. h i−1 That is, we consider the eigenvalue problem for the ˆ a ˆ b ˆ T −1 ˆ ˆ b Pl = Pl I + Hl Rl Hl Pl . (14) (k +1)×(k +1) matrix XbT Xb, whose nonzero eigen- b Finally, going back to the local space represen- values are those of Pl and whose corresponding eigenvectors left-multiplied by Xb are the k eigen- tation, we have vectors u(j) of Pb. a ˆ a b l x¯ l = Q∆x¯ l +x ¯ l . (15) We denote the projection of vectors into the k- dimensional space and the restriction of matrices to 2.4 Ensemble of local analyses the same space by a superscribed circumflex (hat). The ensemble of local analysis fields xa(i) , i = The operator of this projection is l 1, 2, ... , k + 1 is obtained by first finding the (k + 1)  (1) (2) (k) a(i) Q = u | u | · · · | u . (8) local analysis perturbations δxl , a(i) a(i) For instance, for the d-dimensional local back- δxl = Qδxˆ l , (16) b b ground vector xl , the vectorx ˆ l is a k-dimensional then forming the local analysis ensemble column vector given by a(i) a a(i) xl =x ¯ l + δxl . (17) xˆ b = QT xb. (9) l l The local analysis perturbations δxˆ a(i) are a lin- Similarly, for a d × d matrix, such as the local back- ear combination of the local background perturba- b ˆ b tions in the k-dimensional analysis space: ground covariance matrix Pl , the matrix Pl is k × k and is given by ˆ a ˆ b Xl = Xl Yl , (18) ˆ b T b Pl = Q Pl Q. (10) where n o Xˆ a,b = k −1/ 2 δxˆ a,b(1) | δxˆ a,b(2) | · · · | δxˆ a,b(k+1) We also note that, in the internal coordinate system, l l l l ˆ b (19) Pl is diagonal: and   ˆ b (1) (2) (k) 1/ 2 Pl = diag λ , λ , ... , λ , (11)   −1    −1  ˆ bT ˆ b ˆ a ˆ b ˆ b ˆ b Yl = I + Xl Pl Pl − Pl Pl Xl . and so is trivial to invert. (20) This construction of the local analysis perturbations 2.3 Local analysis has the desirable properties that it does not distort o the mean of the analysis ensemble, it correctly rep- Let yl be the vector of current observations within the local region. We solve the Kalman filter equa- resents the analysis uncertainty, and it preserves tion in the local low-dimensional subspaces. Let the smoothness of the background ensemble fields as closely as possible (see Ott et al. 2004 for de- a a b ∆xl = xl − x¯ l . (12) tails). 2.5 Ensemble of global analyses the spectral coefficients of the two-dimensional vor- ticity and divergence are computed from the spec- The components of the most probable global anal- tral coefficients of the two horizontal components of a ysis fieldx ¯ g at the grid point (m, n, o) are ob- the wind vector. (We note that the spectral trans- tained by first selecting the local analysis vector form from grid to spectral space has a smoothing a x¯ l (m, n, o) associated with the local region cen- effect, which that the final analysis used as tered at (m.n, o), then copying the components of model initial condition is smoother than the original a x¯ l (m, n, o) at its central grid point. The same strat- analysis prepared in grid point space.) egy is used to obtain the members of the global Our implementation of the LEKF scheme in- n a(i)o analysis ensemble xg , i = 1, 2, ··· , k + 1. volves both horizontal and vertical localizations of the 192 × 94 × 28 model grid. The horizontal lo- 3. Implementation on the NCEP GFS calization is done on the 192 × 94 longitude-latitude grid. To avoid introducing artificial singularities, ex- 3.1 The forecast model tra latitudes are also added at the poles to make possible the proper definition of the local regions The NCEP GFS is a spectral model, which means on the entire globe. Visually, one may think of our that the model state variables are coefficients of a approach as defining the local regions on a polar spherical harmonic expansion on the globe. We stereographic map projection near the poles. In use a version of the NCEP GFS that was in oper- all experiments reported here, the number of grid ational use at the beginning of 2001. In this ver- points in the zonal and meridional directions are sion the model variables are spectral coefficients equal, while the vertical localization is done by cen- of the two-dimensional vorticity and divergence, vir- tering narrow layers around all 28 model levels. tual temperature, logarithm of the surface pressure, specific humidity, and ozone mixing ratio. The only 3.3 Algorithmic complexity difference between our version and the operational one is in the resolution, which we have reduced to An algorithm is O(f (n)) if, given input data of T62 in the horizontal direction and to 28 level in the length n, the required number of machine instruc- vertical direction. This resolution is well tested inso- tions is bounded by Cf (n) for some constant C when far as many operational forecast products of NCEP n is sufficiently large. This so-called order nota- have been obtained at this resolution for more than tion describes the computational complexity of a a decade. This resolution allows for a large number given algorithm and provides a rough measure of of experiments with the computational resources the computing time on a single processor. For in- available for us. stance, multiplying an n-vector by a scalar is an O(n) procedure. The product of an n × k matrix 3.2 Definition of the local regions with a k-vector is O(nk), insofar as one must form a linear combination of k vectors, each of length n; The nonlinear and physical parameterization terms the classical algorithm for computing the product of in the T62 resolution NCEP GFS are computed on a an n × k matrix with a k × m matrix is O(nkm). 192×94 Gaussian longitude-latitude grid. We utilize Table 1 summarizes the computational com- this grid to implement the LEKF, which is formulated plexity of each step in the LEKF algorithm for a sin- in model grid space. This means that our three- gle local region containing n observations together dimensional global grid has 194 × 92 × 28 points. with d dynamical variables in each of k ensemble Since all variables are defined on the global grid, solutions. In typical applications, d is expected to except for the logarithm of the surface pressure, be one to two orders of magnitude larger than k. the total number of grid point variables is 2 544 768. Therefore, the most expensive steps in the linear (The number of model variables—1 137 024 spec- algebra are the change of coordinates between the tral coefficients—is about 44% of the number of grid model space and the local coordinates (i.e., forming point variables.) Xˆ b and Xa), the computation of the nonzero eigen- b In our experiments, the observed variables are values and associated eigenvectors of Pl , and the the two horizontal components of the wind, the vir- computation of Yl . In the idealized experiments tual temperature, and the surface pressure. We use described here, the observation operator H is triv- this information to obtain an analysis of all grid point ial, multiplication by Hˆ l is simply a gather-scatter variables. Then the spectral transform is applied to operation, and each covariance matrix Rl of local the grid point variables of different types to obtain measurement errors is constant and diagonal. In an analysis in spectral space at each level. Finally, practice, of course, H and Hˆ l are more complex, (Lahey Computer Systems, www.lahey.com) on In- Eq. (1) O(kd) tel Pentium Xeon processors running Red Hat Linux Eq. (4) O(kd) (j) (j) 3 (Red Hat, Inc. www.redhat.com). Because the un- λ , u varies; O(k ) derlying algorithm is iterative, the operation count Xˆ b O(k 2d) associated with DSYEVR is not fixed: the conver- Pˆ a up to O(k 3) gence rate depends on the data, but in general is yo − H x¯ b varies; O(n) here l l expected to be O(k 3). Eq. (13) varies; O(kn) here Eq. (20) varies; O(k 3) The overall efficiency of the LEKF algorithm Xa O(k 2d) also is influenced by the quality of the Fortran intrin- sic function MATMUL, which we use heavily to multi- ply matrices. The level-3 BLAS routines DGEMM and Table 1: The computational complexity of the main steps DGEMV (Dongarra 1990) may be substituted to yield in the LEKF algorithm for each local region: k denotes faster code, depending on the Fortran implementa- the size of the ensemble; d, the dimension of the local tion, though with some loss of clarity and simplicity. atmospheric vector; n, the number of observations in the The total time required to complete the LEKF local region. is proportional to the total number of local re- gions. However, because the assimilation is per- depending on the nature of the data. For con- formed on each region independently, the algorithm ventional observations, such as radiosonde data, is amenable to efficient implementation on paral- H may involve only simple linear interpolation be- lel computer architectures, which substantially re- duces the wall-clock time. tween model grid points, Rl may be banded or even diagonal, and the evaluation of Eq. (13) may still be O(kn). Remotely sensed data, such as satel- 4. Experimental design lite measurements, may involve the evaluation of a We assume that the NCEP GFS provides a perfect nonlinear observation operator and, hence, Eq. (13) representation of the true atmosphere, an approach may be more expensive to compute. However, in frequently called the perfect model hypothesis. Un- either case, if R is approximately constant, then l der this assumption, forecast errors arise and grow R−1 may be precomputed, thus reducing the cost l exclusively due to uncertainties in the initial condi- of evaluating Eq. (13). tions and the sensitivity of the model solutions to Similar comments apply to the evaluation of the these uncertainties. In other words, the model is analysis covariance matrix in Eq. (14). The com- a chaotic system, in the sense that uncertainties in putational complexity may vary from O(k), as in the the initial conditions are more frequently amplified experiments described here, to O(k 3), depending than damped during the forecast phase of the anal- on the structure of R and Hˆ . l l ysis cycle. The role of the data assimilation sys- Our initial implementation of the LEKF is in For- tem, on the one hand, is to use the information con- tran 95, which provides a simple, portable, and effi- tained in the observations to remove the growing cient notation for handling dense matrices. We have component of the errors from the background. On used version 3 of the LAPACK library (Anderson et the other hand, the data assimilation must use the al. 1999) to compute matrix inverses, eigenvalues, information contained in the background to filter the and eigenvectors, because it is numerically robust, observational noise and to spread the information to thoroughly tested, and widely available; most com- unobserved locations. We have designed a series puter vendors provide optimized implementations of of experiments that measure the efficiency of the the Basic Linear Algebra Subroutines (BLAS) (Law- LEKF in all of these three areas (removing growing son et al. 1979; Dongarra 1988 and 1990) upon errors, reducing noise, and spreading information). which the LAPACK library is built. The LAPACK rou- tine DSYEVR implements the algorithm of choice for 4.1 Observations finding all eigenvalues and eigenvectors of a sym- metric k × k matrix. [However, for maximum effi- First, a time series of “true” states, xt (t), was gen- ciency, DSYEVR requires IEEE-754 (IEEE, 1985) in- erated by a 60-day integration of the T62 GFS finity arithmetic to be implemented without trapping model, started from the operational NCEP analy- (Anderson et al. 1999, p. 146)]. This requirement sis at 0000 UTC on 1 January 2000. Then, simu- can be problematic, depending on the processor lated observations were prepared at each grid point and Fortran compiler, although we have had no diffi- by adding zero-mean Gaussian random noise (sim- culty with version 6.1 of the Lahey Fortran compiler ulated observational error) to the true states ev- ery six hours (at 0000 UTC, 0600 UTC, 1200 UTC, 1800 UTC). The of the assumed observational errors is 1 K, 1.1 m/s, and 1 hPa 25 for the virtual temperature, horizontal wind com- ponents, and surface pressure, respectively. The 20 humidity and ozone variables and the physical pa- rameters describing the conditions of the underly- 15 ing surface (e.g., sea surface temperature, albedo, sigma level snow and ice coverage, soil type, etc.) are not ob- 10 served. In most of the experiments, to simulate re- duced observational networks, only subsets of the 5 observations are assimilated. These subsets have been created in a systematic manner, gradually re- 1 2 3 4 5 6 7 8 moving observational locations to obtain sparser number of layers observational data sets.

4.2 Diagnostics FIG. 1: Number of model levels in the local regions (x axis) centered at the different levels (y axis) The accuracy of an analysis field at a given time is measured by the root-mean-square (rms) dis- tance between the analyzed and “true” meteorolog- ical fields. The rms error at a given model level to explore the sensitivity of the scheme to changes is computed by taking the mean over all horizontal in the various assimilation parameters. grid points in the verification area. Time-mean re- sults are then obtained by averaging the rms values over time. When horizontal distributions of time- 5.1 Temporal evolution of errors mean errors are shown, the fields are obtained by averaging the absolute value of the error at each With the rms analysis error rapidly tends to a level grid point. (Following the conventions of numerical that is much smaller than the rms error of the ob- weather prediction, the rms is never taken over both servations. We have found this to be a robust prop- space and time.) erty: it is characteristic of all experiments reported in this paper. While the speed of convergence is 5. Numerical experiments rapid, it is slightly different for the different variables: fastest for the temperature and slowest for the wind Our strategy to validate and tune our implementa- components. Fig. 2 shows an example of the time tion of the LEKF on the NCEP GFS is based on evolution of the rms error for the surface pressure first designing a base experiment, then exploring of the base experiment. In this example, the rms changes in the behavior of the data assimilation error tends to a level that is about 40% of the ob- system under gradual changes to selected param- servational error within a couple of days. While eters of the scheme. In the base experiment, the some slow temporal fluctuations of the error can be surface pressure, horizontal wind components, and observed, the efficient filtering of the observational virtual temperature are observed at 2000 randomly noise by the data assimilation scheme is evident. selected geographical locations, and the parame- Since observations are taken at only 2000 locations ters of the LEKF are the following: the number of (or about 11% of all the grid points), the results also ensemble members is 40, the horizontal size of the indicate that the LEKF efficiently propagates infor- local region is 7 × 7, the depth of the local layers mation to the unobserved locations. varies with altitude (see Fig. 1). The scheme also Since the temporal fluctuation of the errors is applies a uniform 4% multiplicative inflation (Ander- modest, the different configurations of the LEKF can son and Anderson 1999) to the background ensem- be meaningfully compared by the examination of ble to compensate for the variance lost to nonlinear- time-mean results. To ensure that the time means ities and the limited sample size. According to the are not affected by the initial transient, the time av- results of the numerical experiments described be- erages are computed for the last 45 days (the last low, the base configuration is a reasonable, though 180 analysis cycles) of each experiment, i.e., the not optimal, configuration of the LEKF. As the re- first 15 days (60 analysis cycles) are a “spin-up” pe- sults show, this configuration is a good starting point riod whose results are ignored. FIG. 2: Time evolution of the surface pressure rms er- ror for the base configuration (solid line). The rms error FIG. 3: Longitude-averaged time-mean rms error in the of the observations is shown by dashes. temperature analysis for the base configuration (shades). The time-mean of the “true” temperature is also shown (contours). 5.2 Spatial distribution of errors

The geographical distribution of the analysis errors is strongly zonal (Figs. 3 and 4); the largest errors warm air located in the southwest Pacific/northeast are in the tropics and over the polar regions; the Indian region and off the northeast coast of South errors in the mid-latitudes are the smallest. The America. The sources of CAPE are air parcels time-mean error in the temperature analysis is small lifted from the surface. Because the analysis is very compared to the rms observational errors (1 K), accurate near the surface, the temperature flux as- even at the locations of the largest errors. The er- sociated with the ascending warm air parcels may rors for the wind are the largest in the same regions be relatively well analyzed in the subcloud layer. as for the temperature. Despite this similarity, there This indicates that the relatively poor analysis in is an important difference between the errors for the deep convective clouds originates from a poor the two variables: compared to the observational analysis of processes within the clouds themselves. errors, the relative errors are clearly larger for the Also, because the wind analyses are of poorer qual- wind (Figs. 3 and 4). (We recall that the observa- ity than the temperature analysis, it may be con- tional wind error is 1.1 m/s.) siderably more difficult to analyze the momentum Fig. 4 indicates that the analysis of the wind is fluxes in deep convective clouds than the asso- the most difficult in the region of ascending mo- ciated temperature fluxes. These difficulties, en- tions in the Hadley cells. The error quickly de- countered in the regions of deep convection, cannot creases toward the poles in the neighboring regions be explained by the inadequate parameterization of of descending motions. A comparison of Fig. 4 and convection; in our experiments, the true state has Fig. 14.3 of Emanuel (1994) suggests that the er- also been generated by the model. rors are the largest in the layers, where the con- The version of the NCEP GFS that we use em- vective available potential energy (CAPE) is the ploys a modified Arakawa-Schubert scheme (Pan largest. A picture emerges in which the largest er- and Wu 1995) for the parameterization of deep con- rors in the tropical wind analysis are associated with vection. On the one hand, the scheme is sim- deep convective processes. This conclusion was pler than the original Arakawa-Schubert scheme also well supported by an inspection of the geo- (Arakawa and Schubert 1974); it assumes that the graphical (latitude-longitude) distribution of the er- deep convection is associated with one type of rors (not shown). This revealed that the locations of cloud (the deepest cloud) instead of a spectrum of the largest errors are sandwiched between the re- clouds. On the other hand, the scheme corrects gions of easterly trade winds to the north and the one important deficiency of the original scheme: south. These regions are associated with pools of it allows for the transport of momentum by down- model. Our results show that preparing the analysis for the regions of deep convection can be challeng- ing even in the absence of model errors. We later show that in the perfect model set-up, the analysis error can be efficiently reduced in the regions of deep convection by increasing the en- semble size, reducing the depth of the local layers, increasing the number of observations, and using information about the humidity. We also note that since 2001, NCEP has made several important up- grades to the parameterization of deep convection in the operational GFS. We would not be surprised if an implementation of the LEKF on the current op- erational model behaved somewhat differently than that reported here. Another important feature is the good quality of the analysis in the mid-latitude regions. Fig. 3 FIG. 4: Longitude-averaged time-mean rms error in the shows that, in the lower troposphere (below the 50- analysis of the zonal component of the wind for the base kPa level), the errors are practically negligible. This configuration (shades). The time-mean of the “true” circu- is most remarkable, since these are the regions lation in the latitude-height plane is shown by streamlines where baroclinic instabilities—the most energetic (contours with small arrowheads showing the direction of instabilities in the Earth’s atmosphere—convert the the flow). available potential energy to kinetic energy. (The main zones of baroclinic energy conversion can be recognized in Fig. 4 by the on-average upward mo- tions that they generate in the mid-latitudes.) This drafts. Otherwise, the scheme retains the central finding shows the efficiency of the Kalman filter in hypothesis of the Arakawa-Schubert scheme: con- correcting fast-growing background errors associ- vection is essentially a rapid-response mechanism ated with processes initiated by baroclinic instabil- to neutralize the destabilizing effects of such large- ities. The largest, but still modest, analysis errors scale processes as surface fluxes and radiation. in the mid-latitudes occur in and right below the jet More precisely, the scheme assumes that the con- layer. These error patterns are directly connected to sumption of CAPE, by an ever present ensemble the patterns of large errors in the tropics. We have of deep convective clouds, is in a statistical equilib- prepared animations of the error propagation which rium with the CAPE generated by the large scale clearly show that errors in the regions of deep con- processes. vection frequently propagate to the mid-latitudes. One may ask whether our results have any sig- Thus, we conjecture that errors in the mid-latitude nificance in the case where real observations, col- upper troposphere can be reduced by improving the lected in real deep convective clouds, are assimi- analysis of the deep convection in the tropics. lated. As Emanuel (1994) pointed out, the obser- Finally, we note the narrow regions of relatively vational evidence to support the central hypothe- large errors at the top of the atmosphere and over sis of Arakawa and Schubert (1974) is striking. He the polar regions. While the origin of these errors is also pointed out, however, that the entraining plume not clear to us, we suspect that minor flaws in our model, on which the Arakawa-Schubert scheme is implementation of the LEKF may play some role. based, is a poor representation of the individual For instance, we hope that using a reduced grid clouds. Thus, in the regions of deep convection, near the poles would take into account the decreas- we can expect large model errors to occur, which ing distance between grid points and would help to makes a good analysis even more difficult to obtain reduce the errors. Similarly, we may find a better than we have found here. way to prepare the analysis near the upper bound- This example shows the importance of carrying ary of the model atmosphere. While we expect that out experiments under the perfect model hypothe- refinements of our implementation would somewhat sis. Had we started assimilating real observations reduce the errors in these regions, we do not expect first, we might attributed all difficulties to the inad- them to completely eliminate the elevated errors. equate parameterization of deep convection in the The polar regions are similar to the regions of deep convection in that physical parameterizations play an important role, with the important difference that 30 very strong downdrafts characterize the motions in- 25 stead of the updrafts in deep convection. Also, the artificial upper boundary condition (the real atmo- 20 sphere has no upper boundary) is known to gen- erate strong artificial instabilities (Kalnay and Toth 15 1996; Hartman et al. 1997) that may not be effi- sigma level ciently corrected by the data assimilation. Finally, 10 it is important to point out that these errors are still smaller that the observational errors (except for the 5 wind error at the top of the atmosphere), and we have not found indications that these errors prop- 0 0 1000 2000 3000 4000 5000 agate deep into other regions (mid-latitudes and number of variables lower layers of the atmosphere). In what follows, we present a series of exper- iments. In each of these experiments, a parame- FIG. 5: Number of model variables in the local regions ter is selected and gradually altered, starting from (x axis) centered at the different model levels (y axis). the value used in the base experiment. Since we The number of variables is shown for local regions con- have found that the major patterns in the error dis- sisting of 3×3 (black), 5×5 (blue), 7×7 (base experiment, tribution change only a little in the tested parameter red), 9×9 (green) and 11×11 (purple) grid points at each ranges, we describe the effects of changing the se- level. lected parameter by showing average errors over large regions. Errors in the wind analysis are com- ◦ ◦ puted separately for the tropics (30 S–30 N) and sients to settle is plenty in all cases (results are not ◦ ◦ for the two extratropical regions (30 N–90 N and shown). While the rms error of the surface pressure ◦ ◦ 30 S–90 S), but due to the strong similarities be- analysis is essentially the same for all experiments tween the hemispheres, results are not shown for (including the base experiment shown in Fig. 2) the extratropics in the Southern Hemisphere. Also, once the transient dies out, the patch size has a only global averages are shown for the temperature, noticeable impact on the accuracy of the tempera- since the relatively small differences between the ture and wind analyses. The vertical distributions tropics and the extratropics have made a separa- of the time-mean rms error differ with region size tion of the two regions unnecessary. for the temperature over the globe (Fig. 6) and the zonal component of the wind vector in the Northern 5.3 Sensitivity to the size of the local regions Hemisphere extratropics (not shown) and the trop- The size of the local regions is a crucial parameter ics (Fig. 7), albeit not dramatically. The results are of the LEKF scheme. On the one hand, the local the best for the 5 × 5 horizontal regions, providing region must be sufficiently large to allow nearby ob- uniformly good results over the different variables servations, which contain useful information, to af- and regions. In the extra-tropics, the 7 × 7 local re- fect the analysis at the center of the region. On the gions provide good performance similar to the 5 × 5 other hand, the region must be sufficiently small to local regions, while in the tropics, the 5 × 5 local re- permit an efficient filtering of statistical fluctuations, gions have a clear advantage over the 7×7 and the which occur due to the relatively small ensemble similarly well-performing 3 × 3 regions. Increasing size. (The computational cost increases as much as the size of the horizontal regions to 9×9, and espe- cubically with the number of ensemble members.) cially to 11 × 11, clearly degrades the performance The best way to find a nearly optimal region size is of the scheme. through numerical experimentation. In summary, we conclude that the rms error of First we vary the horizontal size of the local re- the analyses is about 30% of the observational er- gions, using 3 × 3, 5 × 5, 7 × 7, 9 × 9, and 11 × 11 rors, when the ensemble has k = 40 members and patches of grid points. The dimension d of the lo- the patch-size is selected from a reasonable range cal vectors for the different-sized regions is shown (from 3 × 3 to 7 × 7 horizontal grid points at the in Fig. 5. The horizontal size of the local regions number of vertical levels shown in Fig. 1. has no significant effect on the speed of the conver- It is reasonable to assume that when the en- gence, and the 15 days that we allow for the tran- semble size is increased, detection of the (typically FIG. 8: Same as Fig. 7, but for an 80-member ensem- ble. [The results are virtually the same for the 5 × 5 (blue) FIG. 6: The time mean of the rms error of the tem- and 7 × 7 (base experiment, red) regions and are not perature analysis over the globe (x axis) as a function of shown for the 9 × 9 (green) and 11 × 11 (purple) horizon- height, using pressure as the vertical coordinate (y axis). tal regions.] The color scheme is the same as in Fig. 5. The order of the errors, from small to begin, is 5 × 5 (blue), 7 × 7 (base experiment, red), 3 × 3 (black), 9 × 9 (green) and 11 × 11 (purple). smaller) longer-distance correlations in the back- ground uncertainties becomes less sensitive to sta- tistical fluctuations. In other words, we can expect that, for a larger ensemble, the optimal region size also increases and the better utilization of more ob- served information leads to a reduction of the anal- ysis errors. To assess the effect of ensemble size on the optimal region size and the analysis errors, experiments with k = 80 ensemble members have also been carried out. We find that the increased ensemble size has the largest positive effect on the configuration using 7×7 horizontal grid points. This configuration now breaks even with the configura- tion using 5 × 5 horizontal grid points in all regions and for all variables. This occurs as the perfor- mance of the latter configuration is only slightly af- fected by the increased ensemble size. These re- sults indicate that a 40-member ensemble suffices to resolve the analysis uncertainty in the 5 × 5 local horizontal regions. Since the overall improvement from increase of the ensemble size from 40 to 80 is very modest for the extratropics, the results are illustrated only for the tropics (Fig. 8). FIG. 7: The time mean of the rms error of the zonal wind component analysis in the tropics (x axis) as a func- 5.4 Sensitivity to the density of observations tion of height, using pressure as the vertical coordinate (y axis). The color scheme is the same as in Fig. 5. The Earlier experiments with simple models show that rms error of the observations is shown by dashes. ensemble Kalman filters have a growing advantage over 3D-Var-type schemes, which use static esti- mates of the background covariance matrix, as the density of the observations decreases (e.g., Hamill FIG. 9: The time mean of the rms error of the tem- FIG. 10: The time mean of the rms error of the zonal perature analysis over the globe (x axis) as function of wind component analysis in the tropics (x axis) as a func- height using pressure as vertical coordinate (y axis). The tion of height using pressure as the vertical coordinate analysis error is shown for observing networks consist- (y axis). The color scheme is the same as in Fig. 5.4. ing of 18 048 (blue), 2000 (base experiment, red), 1000 (green) and 500 (purple) observational locations. This corresponds to observing all grid points (blue), and 11% (red), 5.5% (green), and 2% (purple) of all grid points. 5.5 Timing results The rms error of the observations is shown by dashes. The timing results described here are from our ini- tial implementation of the LEKF, which is on a rela- tively modest Beowulf cluster consisting of 25 dual- processor nodes, each with 2 GB of random ac- cess memory and connected by a 1-gigabit Eth- ernet; each processor is a 2.8-GHz Intel Pentium et al. 2001; Ott et al 2004). More precisely, while Xeon with hyperthreading disabled. Most runs use the performance of the 3D-Var-type schemes de- 40 processors on the cluster. grades dramatically with decreasing observational Typically, one complete cycle of the algorithm density, the accuracy of the Kalman filter is only takes 15 minutes of wall-clock time for an ensem- slightly affected until the density of observations ble of 40 solutions when observations are available reaches a critically-low value. As the density of at every model grid point using the GFS at T62/L28 observations decreases, the correct estimation of resolution and vertical localization, as described be- the flow-dependent covariance becomes ever more low. This computation involves 192 × 94 × 28 = reliant on locations to which information has been 505 344 local regions and slightly more than 1.5 propagated from elsewhere.To test whether our im- million observations; each local region is a cube plementation of the LEKF on the NCEP GFS re- of 7 × 7 × v model grid points, where v = 1, 3, tains this important property, further experiments 5 or 7, depending on altitude. The time includes with differing number of observations were also car- that spent computing the transforms from spectral ried out. The results are compared for four different space to physical space and back (which could be observational networks: observing all, 2000, 1000, parallelized but which we have implemented only and 500 locations around the globe. While the ob- on a single processor); the i/o and network over- servational density has only a modest impact on the head to transport the appropriate model grid and quality of the temperature analysis (Fig. 9), the ac- observation data to each processor; one step of curacy of the wind analysis, especially in the tropics, the LEKF algorithm to each local region; and 40 degrades more dramatically as the density of the six-hour forecasts from the resulting analysis. Ex- observations is reduced (Fig. 10). Most importantly, cluding the spectral transforms and forecasts, the Fig. 10 shows that increasing the number of obser- wall-clock time is about 505 seconds for the above vations is a very efficient way to reduce the analysis parameters. The average time needed to process errors in the tropics (in regions of deep convection). one local region, where the ensemble size is 40 rate and computationally efficient data assimilation parameters wall-clock LEKF step patch mean scheme. Based on the results presented here, we −3 k + 1, N time (min) only (sec) (10 sec) believe that the LEKF is an operationally attainable. 40, all 15 505 31 We hope to further test the potentials of the LEKF 40, 2000 14 447 28 for operational purposes by an implementation on 80, all 45 1973 122 a state-of-the-art, high resolution, operational re- 80, 2000 42 1682 104 gional model and by assimilating real observations into both the global and the regional models. A rela- Table 2: Time needed to run the LEKF algorithm. tively quick implementation is made possible by the important feature of the LEKF algorithm, that it is in- dependent of the details of a given weather model. and the patch size is 490, is about 31 millisec- The code required to interface with the GFS model onds. Table 2 shows timing results for local regions is localized into a module called “Grid Manager,” that consist of 7 × 7 × v cubes (v = 1, 3) as de- which handles the interface between the spectral scribed above, with either 40 or 80 ensemble mem- transforms and the physical grid used for the data bers. (For this calculations the v=5,7 layers were assimilation. The entire software suite, including replaced by layers of v=3.) The observing network code to implement the LEKF algorithm, spectral consists of observations of temperature and wind transforms, and diagnostics, comprises about 8500 speed at all 28 vertical levels, plus the surface pres- lines of Fortran 95 code (including extensive com- sure, at each of N points. The notation N = all ments). About half of the total is contained in the refers to the case where observations are available Grid Manager and the spectral transforms; only this at each model grid point (1.5 million observations portion of the code needs to be rewritten to apply in total); N = 2000 refers to an observing network the algorithm to a different forecast model. that consists of 2000 randomly-chosen longitude- latitude model grid points at which the observations 7. Acknowledgments are assumed to exist at each vertical level (i.e., a to- tal of 170 000 observations). The wall-clock time in- This work was supported by the Army Research Of- cludes the total time needed to perform all spectral fice, a James S. McDonnell 21st Century Research transforms, the LEKF algorithm, and k + 1 six-hour Award, by the Office of Naval Research (Physics), forecasts from the resulting analysis. The column and by the National Science Foundation (Grants labeled “LEKF only” refers to the maximum amount #0104087 and PHYS 0098632). of time that any given processor spends performing only the data assimilation step after the model grid 8. References is distributed more-or-less evenly across the clus- ter; the last column shows the average time spent Anderson, E. et al., 1999: LAPACK Users’ Guide, processing a single local region. third ed. Society for Industrial and Applied Mathe- These results suggest that, for a given observ- matics, Philadelphia. ing network and patch size, the overall time re- quired to perform the LEKF assimilation step grows Anderson, J. L., 2001: An ensemble adjustment fil- roughly quadratically with the number of ensemble ter for data assimilation. Mon. Wea. Rev., 129, solutions and is relatively insensitive to the amount 2884-2903. of data to be assimilated. In practice, of course, the work required to decode observations from data Anderson, J. L., and S. L. Anderson, 1999: A Monte files and evaluate the observation operators may Carlo implementation of the nonlinear filtering prob- be substantially greater than what is needed here. lem to produce ensemble assimilations and fore- Nevertheless, we are optimistic that, with further casts. Mon. Wea. Rev., 127, 2741-2758. tuning and a somewhat larger computer, the LEKF IEEE, 1985: IEEE Standard for Binary Floating- data assimilation algorithm can be performed within Point Arithmetic. ANSI/IEEE Std. 754-1985. Ameri- the time constraints of a typical operational forecast can National Standards Institute. center. Anderson, E. et al., 1999: LAPACK Users’ Guide, 6. Conclusion third ed. Society for Industrial and Applied Mathe- matics, Philadelphia. We have demonstrated, by an implementation on the NCEP GFS, that the LEKF is a highly accu- Arakawa, A., and W. H. Schubert, 1974: Interaction of a cumulus cloud ensemble with the large-scale Reading, United Kingdom, ECMWF, 69-82. environment. Part I. J. Atmos. Sci., 31, 674–701. Keppenne, C., and H. Rienecker, 2002: Initial test- Bishop, C. H., B. J. Etherton, and S. Majumdar, ing of a massively parallel ensemble Kalman Filter 2001: Adaptive sampling with the Ensemble Trans- with the Poseidon Isopycnal Ocean General Circu- form Kalman Filter. Part I: Theoretical aspects. lation Model. Mon. Wea. Rev., 130, 2951-2965. Mon. Wea. Rev., 129, 420-436. Lahey Computer Systems. www.lahey.com. Dongarra, J. J. et al., 1988: An extended set Lawson, C. L. et al.,1979: Basic Linear Algebra of FORTRAN Basic Linear Algebra Subprograms. Subprograms for FORTRAN usage. ACM Trans. ACM Trans. Math. Soft., 14, 1–17. Math. Soft. 5, 308–323. Dongarra, J. J. et al.,1990: A set of Level 3 Basic Lorenz, E. N., 1996: Predictability: A problem partly Linear Algebra Subprograms. ACM Trans. Math. solved. Proc. Seminar on Predictability, Vol. 1, Soft. 16, 1–17. ECMWF, Reading, Berkshire, UK, 1-18. Dowell, D. C., F. Zhang, L. J. Wicker, C. Snyder, and Ott, E., B. R. Hunt, I. Szunyogh, M. Corazza, E. N. A. Crook, 2004: Wind and thermodynamic re- Kalnay, D. J. Patil, J. A. Yorke, A. V. Zimin, and E. trievals in the 17 May 1981 Arcadia, Oklahoma, su- Kostelich, 2002: Exploiting local low dimensionality percell: Ensemble Kalman filter experiments. Mon. of the atmospheric dynamics for efficient Kalman fil- Wea. Rev., 132, (in press) tering. http://arxiv.org/abs/physics/0203058. Emanuel, K. A., 1994: Atmospheric convection. Ott, E., B. R. Hunt, I. Szunyogh, A. V. Zimin, E. J. Oxford University Press, New York. Kostelich et al., 2004: A local ensemble Kalman fil- Evensen, G. 2003: The ensemble Kalman filter: ter for atmospheric data assimilation. Tellus, 56A, Theoretical formulation and practical implementa- (in press). tion. Ocean Dynamics, 53, 343–367. Pan, H.-L. and W.-S. Wu, 1995: Implementing Hamill, T. M., J. Whitaker, and C. Snyder, 2001: a Mass Flux Convection Package for the NMC Distance-dependent filtering of background error Medium-Range Forecast model. NMC Office Note covariance estimates in an Ensemble Kalman Fil- 409. Available from NCEP, 5200 Auth Road, Wash- ter. Mon. Wea. Rev., 129, 2776-2790. ington, 20233. Hartman, D. L., T. N. Palmer, and R. Buizza, 1995: Sauer, T., B. R. Hunt, J. A. Yorke, A. V. Zimin, and Finite-time instabilities of lower-stratospheric flow. E. Ott et al., 2004: 4D Ensemble Kalman filter- J. Atmos. Sci., 53, 2129-2143. ing for assimilation of asynchronous observations. Preprints of the 20th Conference on Weather Anal- Houtekamer, P. L., and H. L. Mitchell, 2001: A se- ysis and Forecasting/16th Conference on Numeri- quential ensemble Kalman Filter for atmospheric cal Weather Prediction. Seattle, WA, USA, January data assimilation. Mon. Wea. Rev., 129, 796-811. 11–15, 2004. Houtekamer, P. L., H. L. Mitchell, G. Pellerin, Snyder, C. and F. Zhang, 2003: Assimilation of sim- M. Buehner, M. Charron, L. Spacek, and B. ulated Doppler radar observations with an Ensem- Hansen, 2003: Atmospheric data assimilation with ble Kalman filter. Mon. Wea. Rev., 131, 1663-1677. the ensemble Kalman filter: Results with real observations. Submitted. Avaibale from pe- Szunyogh, I., E. J. Kostelich, G. Gyarmati, B. Hunt, [email protected]. and E. Kalnay et al., 2004: A Local Ensemble Kalman Filter for the NCEP GFS. Preprints of the Hunt, B. R., E. Kalnay, E. J. Kostelich, E. Ott, and 20th Conference on Weather Analysis and Fore- D. J. Patil et al., 2004: Four-dimensional ensemble casting/16th Conference on Numerical Weather Kalman filtering. Tellus, 56A, (in press). Prediction. Seattle, WA, USA, January 11–15, 2004. Kalnay, E., 2003: Atmospheric modeling, data as- similation, and predictability. Cambridge University Tippett, M. K., J. L. Anderson, C. H. Bishop, T. M. Press, Cambridge, 341 pp. Hamill, and J. S. Whitaker, 2003: Ensemble square- root filters. Mon. Wea. Rev., 131, 1485-1490. Kalnay, E. and Z. Toth, 1996: The breeding method. Proc. of the ECMWF Seminar on Predictability, Whitaker, J. S., and T. H. Hamill, 2002: Ensemble Data Assimilation without perturbed observations. Mon. Wea. Rev., 130, 1913-1924. Whitaker, J. S., G. P. Compo, X. Wei, and T. M. Hamill, 2003: Reanalysis without radioson- des using ensemble data assimilation. Mon. Wea. Rev., submitted. Available from jef- [email protected]. Zhang, F., C. Snyder, and J. Sun, 2004: Impacts of initial estimate and observation availability on convective-scale data assimilation with an Ensem- ble Kalman filter. Mon. Wea. Rev., 132, in press.