ArticlePDF Available

Efficient Implementation of the Ensemble Kalman Filter

Authors:

Abstract

We present several methods for the efficient implementation of the Ensemble Kalman Filter (EnKF) of Evensen. It is shown that the EnKF can be implemented without access to the observation matrix, and only an observation function is needed; this greatly simplifies software design. New implementations of the EnKF formulas are proposed, with linear computational complexity in the number of data points. These implementations are possible when the data covariance matrix is easy to decompose, such as a diagonal or a banded matrix, or given in a factored form as sample covariance. Unlike previous methods, our method for the former case uses Choleski decomposition on a small matrix from the Sherman-Morrison-Woodbury formula instead of SVD on a large matrix, and our method in the latter case does not impose any constraints on data randomization. One version of the EnKF formulas was implemented in a distributed parallel environment, using SCALAPACK and MPI.
University of Colorado at Denver and Health Sciences Center
Efficient Implementation of
the Ensemble Kalman Filter
Jan Mandel
May 2006 UCDHSC/CCM Report No. 231
CENTER FOR COMPUTATIONAL MATHEMATICS REPORTS
EFFICIENT IMPLEMENTATION OF
THE ENSEMBLE KALMAN FILTER
JAN MANDEL
Abstract. We present several methods for the efficient implementation of the Ensemble Kalman
Filter (EnKF) of Evensen. It is shown that the EnKF can be implemented without access to the
observation matrix, and only an observation function is needed; this greatly simplifies software design.
New implementations of the EnKF formulas are proposed, with linear computational complexity in
the number of data points. These implementations are possible when the data covariance matrix is
easy to decompose, such as a diagonal or a banded matrix, or given in a factored form as sample
covariance. Unlike previous methods, our method for the former case uses Choleski decomposition on
a small matrix from the Sherman-Morrison-Woodbury formula instead of SVD on a large matrix, and
our method in the latter case does not impose any constraints on data randomization. One version
of the EnKF formulas was implemented in a distributed parallel environment, using SCALAPACK
and MPI.
1. Introduction. The Ensemble Kalman Filter (EnKF) is a Monte-Carlo
implementation of the Bayesian update problem: Given a probability distribution
of the modeled system (the prior, called often the forecast in geosciences) and data
likelihood, the Bayes theorem is used to to obtain the probability distribution with
the data likelihood taken into account (the posterior or the analysis). The Bayesian
update is combined with advancing the model in time, with the data incorporated
from time to time. The original Kalman Filter [14] relies on the assumption that
the probability distributions are Gaussian, and provided algebraic formulas for the
change of the mean and covariance by the Bayesian update, and a formula for
advancing the covariance matrix in time provided the system is linear. However,
this is not possible computationally for high dimensional systems. For this reasons,
EnKFs were developed [7, 12]. EnKFs represent the distribution of the system state
using a random sample, called an ensemble, and replace the covariance matrix by the
sample covariance of the ensemble. One advantage of EnKFs is that advancing the
probability distribution in time is achieved by simply advancing each member of the
ensemble. EnKFs, however, still rely on the Gaussian assumption, though they are
of course used in practice for nonlinear problems, where the Gaussian assumption is
not satisfied. Related filters attempting to relax the Gaussian assumption in EnKF
include [3, 4, 15, 17].
This paper is focused on the Bayesian update step for the EnKF version from
[6, 8]. This filter involves randomization of data. For filters without randomization
of data, see [2, 9, 16].
The paper is organized as follows. In Sec. 2, we state the Bayesian update formulas
in EnKF. In Section 3, we show how to evaluate formulas without an observation
matrix, involving observation functions. Sec. 4 contains the discussion of several
implementations of the EnKF and their computational complexity. Finally, Sec. 5
briefly reports on the experience from a distributed parallel implementation.
2. Notation and the EnKF Formulas. We review the EnKF formulas
following [6, 8], with only one minor difference. The forecast ensemble consists of
Nmembers, which are state vectors of dimension n. The ensemble can be written as
Center for Computational Mathematics, University of Colorado at Denver and Health Sciences
Center, Denver, CO 80217-3364
1
the Nby nmatrix
Xf= [x1,...,xN] = [xi]
are the forecast ensemble states. The ensemble mean is
E(X) = 1
N
N
X
k=1
xk.
and the ensemble covariance matrix is
C=AAT
N1
where
A=XE(X) = X1
N(XeN×1)e1×N,
and edenotes the matrix of all ones of the indicated size.
The data is given as a measurement vector dsize mand error covariance matrix
Rsize mby m. Then the measurement matrix Dsize mby Nis defined by
D= [d1, d2, ..., dN], dj=d+vj, vjN(0, R)
where vjare independent random perturbations.
The analysis ensemble is then given by
Xa=X+CH TH CHT+R1(DHX),(2.1)
cf., [8, eq. (20)].
The difference between (2.1) and [8, eq. (20)] is that we use the covariance matrix
Rof the measurement error rather than the sample covariance DDT/(N1) of the
randomized data. Because Ris always positive definite, there is no difficulty with the
inverse in (2.1), unlike [8, eq. (20)].
The analysis formula (2.1) can be rewritten similarly as in [8, eq. (54)] as
Xa=X+1
N1A(HA)TP1(DHX ) (2.2)
where
HA =HX 1
N((HX)eN×1)e1×N,(2.3)
P=1
N1HA (HA)T+R. (2.4)
3. Observation Matrix-Free Method. Clearly, the matrix Hin (2.2) (2.4)
is needed only in the matrix vector product HX, which needs to be computed only
once. However, it is very inconvenient to operate with the matrix Hexplicitly for
several reasons:
1. An observation function h(x) that creates synthetic data from the state can
be quite complicated. Even when the observation function is affine, thus of
the form
h(x) = Hx +f, (3.1)
creating the matrix Hand the vector fis an additional effort, which is
typically much harder than programming the observation function itself.
2
2. Computing the matrix-vector product HX takes computational resources.
3. The matrix His typically sparse and it can be very large. E.g., in the
quite common case when every measurement coincides with the value of
a state variable, the matrix will have just one one in every row and the
rest of its entries are zeros. If the measurement is an image, the number of
measurements can be very large, but each pixel in the image is interpolated
from just a few entries of the state vector. Assimilation of image data
into a geophysical model may easily require m106and n106, which
makes manipulation of the matrix Hstored as full impossible on current
computers. So, the matrix Hmust be stored as sparse, which is an additional
complication.
Fortunately, creating the matrix Hexplicitly is not needed. Assume that we only
have access to the evaluation of h(x) in (3.1) but not to the values of Hor f. To
compute the data residual, note that
dHx =d+f(Hx +f) = e
dh(x),
where
e
d=d+f
is the data actually given. To compute HA, note that
[HA]i=HxiH1
n
n
X
j=1
xj
= (Hxi+f)1
N
N
X
j=1
(Hxj+f)
=h(xi)1
N
N
X
j=1
h(xj).
Consequently, the ensemble update can be computed by evaluating the observation
function hon each ensemble member once.
4. Computational Complexity. All operations in (2.2) (2.4) are evaluations
of the observation function h(x), cf., Sec. 3, and matrix-matrix and matrix-vector
operations, which can be implemented efficiently by calls to the LAPACK, BLAS,
and SCALAPACK libraries [1, 5]. These libraries contain routines for operations
including Choleski decomposition of a symmetric positive definite matrix (CHOL)
LLT=A, and matrix multiply (GEMM), A=αA +βB(T)C(T), where (T) is either
transpose or nothing, for rank update. Recall that mis the number of data points, n
is the number degrees of freedom, Nis the ensemble size, so Xis nby N,HX and
HA are mby N,Ris mby m. Assume that evaluation of h(x) costs O(m), and that
multiplication of matrices of size n1by n2and n2by n3costs O(n1n2n3).
4.1. Reference Implementation. Straightforward implementation of the
formulas (2.2) (2.4) leads to the following algorithm:
3
computation operation size cost
HX N times h(x)m×N O(mN )
z= (HX)eN×1matrix multiply m×N×1O(mN)
HA =HX 1
Nze1×Nmatrix multiply m×1×N O(mN)
Y=DHX matrix add m×N O(mN )
P=R+1
N1HA (HA)Tmatrix multiply m×N×m O(m2N)
LLT=PCholeski m O(m3)
M=P1Ysolution m×m×N O(m2N)
Z= (HA)TMmatrix multiply N×m×N O(mN 2)
Xa=X+1
N1AZ matrix multiply n×N×N O(nN 2)
(4.1)
The total computational complexity of the algorithm (4.1) is
Om3+m2N+mN2+nN2.
So, this method is suitable for a large number of degrees of freedom n, but not for a
large number of observations m.
4.2. Large Number of Data Points. In practice, the number of data points
mis often large, while the data error covariance matrix Ris often diagonal (when the
data errors are uncorrelated), nearly diagonal, or easy to decompose. This happens,
e.g., in the assimilation of images, or in regularized EnKF where the gradient of fields
in the state is assimilated as an additional artificial observation [13]. In this case,
the following algorithm, which has only linear complexity in m, provides a significant
advantage. Assume that multiplication by R1is dominated by other costs. Using
the Sherman-Morrison-Woodbury formula [11]
(R+UV T)1=R1R1U(I+VTR1U)1VTR1,
with
U=1
N1HA, V =HA,
we have
P1=R+1
N1HA (HA)T1
=R1"I1
N1(HA)I+ (HA)TR11
N1(HA)1
(HA)TR1#.
4
The computation (4.1) then becomes
computation operation size cost
HX N times h(x)m×N O(mN )
z= (HX)eN×1matrix multiply m×N×1O(mN)
HA =HX 1
Nze1×Nmatrix multiply m×1×N O(mN)
Y=DHX matrix add m×N O(mN )
Q=I+ (HA)TR11
N1(HA) matrix multiply N×m×M O(mN 2)
LLT=QCholeski N O(N3)
Z= (HA)TR1Ymatrix multiply N×m×N O(mN 2)
W=Q1Zsolution N×N O N3
M=R1hI1
N1(HA)Wimatrix multiply m×N×N O(mN 2)
Z= (HA)TMmatrix multiply N×m×N O(mN 2)
Xa=X+1
N1AZ matrix multiply n×N×N O(nN 2)
(4.2)
This gives overall complexity ON3+mN2+nN2, which is suitable for large nand
large m.
4.3. Square Root Alternative. Decompose first R=SST, e.g., by Choleski
decomposition, and multiplication by S1is cheap to compute. Let e
U=S1U,
e
V=S1V. Then the Sherman-Morrison-Woodbury formula becomes
(SST+U V T)1=STIe
UI+e
VTe
U1e
VS1,
which gives, again with R=SST,B=S1H A
P1=ST"I1
N1BI+1
N1BTB1
BT#S1
and one proceeds just as in Sec. 4.2. The asymptotic complexity is same as in Sec.
4.2, but the formulas involve symmetric products of matrices, which is numerically
more stable and allows to save memory by storing just one triangle.
4.4. SVD Method for Full Data Error Covariance. Assume again that
R=SSTand multiplication by S1is cheap to compute. Write
P=1
N1HA (HA)T+R=S1
N1S1HA (HA)TST+IST
and use the singular value decomposition (SVD) [10]
S1HA =UΣVT,
where Uand Vare orthogonal square matrices, and Σ = diagm×N(σ1, ..., σk) is the
diagonal matrix size mby Nwith the singular values σ1, ..., σk,k= min {m, N }, on
the diagonal; then, using the orthogonality relations VTV=I,UUT=I, we have
1
N1S1HA (HA)TST+I=1
N1UΣUT+I=UΣ2
N1+IUT
5
thus
P1=R1/2UTdiag
1
σ2
i
N+1 + 1
UR1/2
The complexity is again O(mN2),assuming that m > N and that computation of
SVD of matrix size mby Ncosts O(mN min(m, N)).
4.5. SVD and Eigenvalue Methods for Sample Data Error Covariance.
It should be noted that the use of SVD in Sec. 4.4 is different than in [8], where
spectral methods and SVD were advocated as a device to overcome the singularity of
the data sample covariance matrix R, given as R=SST/(N1) with Ssize m×N.
In that case, [8, eq. (56)] suggests the eigenvalue decomposition for the matrix
P=1
N1HA (HA)T+SST=ZΛZT,(4.3)
which is size mby mbut there are only Nnonzero eigenvalues, i.e., diagonal entries of
Λ. If the matrix Pis created explicitly, the cost of the decomposition (4.3) is between
Om2Nand Om3.
However, we note that in this case, Pcan be written as
P=1
N1F F T, F = [HA, S],(4.4)
where Fhas size mby 2N. Therefore, applying the SVD decomposition to the matrix
W, we obtain W=UΣVT, and
P1= (N1) UΣ2UT.(4.5)
The matrix Uis size mby mbut it has only 2Nonzero columns, and SVD routines
actually return the mby 2Nsubmatrix, at the cost OmN 2. The resulting
algorithm, using the multiplication by factored P1from (4.5) in (2.1), has again
the cost ON3+mN2+nN2. [8, eq. (57)] discusses a method using SVD, with the
same asymptotic cost, but requiring the data perturbation selected in a special way.
No such constraint is needed here.
4.6. Iterative Methods. The linear system P M =Ycan be solved by
conjugate gradients for Nright-hand sides. However, each iteration costs OmN2,
so iterative methods do not seem to be competitive.
5. Distributed Parallel Implementation. The method described in Sec. 4.2
was implemented in a distributed parallel environment using MPI and SCALAPACK.
EnKF is naturally parallel: each ensemble member can be advanced in time
independently. The linear algebra in the Bayesian update step links the ensemble
members together. The ensemble matrix Xis then naturally distributed so that
each process owns a block of columns. However, such distribution is a bottleneck to
parallelism: SCALAPACK requires that all matrices involved in an operation must
be distributed on the same processor grid (though possibly with different block sizes),
and, for best performance, the processor grid should be close to a square. 1D processor
grids tend to be particularly inefficient. Therefore, the ensemble matrix must be
redistributed before the matrix linear algebra operations.
6
6. Acknowledgements. Section 4.4 is based on a discussion with Andrew
Knyazev and Craig Johns. The author would like to thank Craig Johns and Mingshi
Chen for useful discussions about the algebra of EnKF, and Jonathan Beezley for
reading this paper. Thanks are also due to Jonathan Beezley, Craig Douglas, Deng Li,
and Adam Zornes for contributing to an object oriented interface to SCALAPACK,
to Craig Douglas and Wei Li for assistance with MPI wrappers, and to Jonathan
Beezley for assistance with a parallel distributed implementation of EnKF on top
of SCALAPACK. This research was supported by the National Science Foundation
under the grant CNS-0325314. Computer time on IBM BlueGene/L supercomputer
was provided by NSF MRI Grants CNS-0421498, CNS-0420873, and CNS-0420985,
NSF sponsorship of the National Center for Atmospheric Research, the University of
Colorado, and a grant from the IBM Shared University Research (SUR) program.
REFERENCES
[1] E. Anderson, Z. Bai, C. Bischof, S. Blackford, J. Demmel, J. Dongarra, J. Du Croz,
A. Greenbaum, S. Hammarling, A. McKenney, and D. Sorensen,LAPACK Users’
Guide, Society for Industrial and Applied Mathematics, Philadelphia, PA, third ed., 1999.
[2] J. L. Anderson,An ensemble adjustment Kalman filter for data assimilation, Monthly
Weather Review, 129 (1999), pp. 2884–2903.
[3] J. L. Anderson and S. L. Anderson,A Monte Carlo implementation of the nonlinear filtering
problem to produce ensemble assimilations and forecasts, Monthly Weather Review, 127
(1999), pp. 2741–2758.
[4] T. Bengtsson, C. Snyder, and D. Nychka,Toward a nonlinear ensemble filter for high
dimensional systems, Journal of Geophysical Research - Atmospheres, 108(D24) (2003),
pp. STS 2–1–10.
[5] L. S. Blackford, J. Choi, A. Cleary, E. D’Azevedo, J. Demmel, I. Dhillon, J. Dongarra,
S. Hammarling, G. Henry, A. Petitet, K. Stanley, D. Walker, and R. C. Whaley,
ScaLAPACK Users’ Guide, Society for Industrial and Applied Mathematics, Philadelphia,
PA, 1997.
[6] G. Burgers, P. J. van Leeuwen, and G. Evensen,Analysis scheme in the ensemble Kalman
filter, Monthly Weather Review, 126 (1998), pp. 1719–1724.
[7] G. Evensen,Sequential data assimilation with nonlinear quasi-geostrophic model using Monte
Carlo methods to forecast error statistics, Journal of Geophysical Research, 99 (C5) (1994),
pp. 143–162.
[8] G. Evensen,The ensemble Kalman filter: Theoretical formulation and practical
implementation, Ocean Dynamics, 53 (2003), pp. 343–367.
[9] ,Sampling strategies and square root analysis schemes for the EnKF, Ocean Dynamics,
(2004), pp. 539–560.
[10] G. H. Golub and C. F. V. Loan,Matrix Computations, Johns Hopkins Univ. Press, 1989.
Second Edition.
[11] W. W. Hager,Updating the inverse of a matrix, SIAM Rev., 31 (1989), pp. 221–239.
[12] P. Houtekamer and H. L. Mitchell,Data assimilation using an ensemble Kalman filter
technique, Monthly Weather Review, 126 (1998), pp. 796–811.
[13] C. J. Johns and J. Mandel,A two-stage ensemble Kalman filter for smooth data assimilation.
Environmental and Ecological Statistics. Conference on New Developments of Statistical
Analysis in Wildlife, Fisheries, and Ecological Research, Oct 13-16, 2004, Columbia, MI,
in print.
[14] R. E. Kalman,A new approach to linear filtering and prediction problems, Transactions of the
ASME Journal of Basic Engineering, Series D, 82 (1960), pp. 35–45.
[15] J. Mandel and J. D. Beezley,Predictor-corrector ensemble filters for the as-
similation of sparse data into high dimensional nonlinear systems. CCM Re-
port 232, University of Colorado at Denver and Health Sciences Center, 2006;
http://www.math.cudenver.edu/ccm/reports, May 2006.
[16] M. K. Tippett, J. L. Anderson, C. H. Bishop, T. M. Hamill, and J. S. Whitaker,Ensemble
square root filters, Monthly Weather Review, 131 (2003), pp. 1485–1490.
[17] P. van Leeuwen,A variance-minimizing filter for large-scale applications, Monthly Weather
Review, 131 (2003), pp. 2071–2084.
7
CENTER FOR COMPUTATIONAL MATHEMATICS REPORTS
University of Colorado at Denver and Health Sciences Center Fax: (303) 556-8550
P.O. Box 173364, Campus Box 170 Phone: (303) 556-8442
Denver, CO 80217-3364 http://www-math.cudenver.edu/ccm
220 Andrew V. Knyazev and Merico E. Argentati, “Implementation of a Preconditioned
Eigensolver Using Hypre.” April 2005.
221 Craig Johns and Jan Mandel, “A Two-Stage Ensemble Kalman Filter for Smooth Data
Assimilation.” April 2005
222 I. Lashuk, M. E. Argentati, E. Ovchinnikov and A. V. Knyazev, “Preconditioned Eigen-
solver LOBPCG in hypre and PETSc.” September 2005.
223 A. V. Knyazev and M. E. Argentati, “Majorization for Changes in Angles Between
Subspaces, Ritz values, and graph Laplacian spectra.” September 2005.
224 D. W. Dean, “A Semimartingale Approach For Modeling Multiphase Flow In Hetero-
geneous Porous Media.” November 2005.
225 D.W. Dean, T.H. Illangasekare and T.F. Russell, “A Stochastic Differential Equation
Approach For Modeling Of NAPL Flow In Heterogeneous Porous Media.” November
2005.
226 J. Mandel and B. Soused´ık, “Adaptive Coarse Space Selection in the BDDC and
the FETI-DP Iterative Substructuring Methods: Optimal Face Degrees of Freedom.”
November 2005.
227 J. Mandel and B. Soused´ık, “Adaptive Selection of Face Coarse Degrees of Freedom in
the BDDC and the FETI-DP Iterative Substructuring Methods.” November 2005.
228 Craig C. Douglas, Jonathan D. Beezley, Janice Coen, Deng Li, Wei Li, Alan K. Man-
del, Jan Mandel, Guan Qin, and Anthony Vodacek, “Demonstrating the Validity of a
Wildfire DDDAS.” February 2006.
229 Lynn Schreyer Bennethum, “Flow and Deformation: Understanding the Assumptions
and Thermodynamics, Updated.” February 2006.
230 Lynn Schreyer Bennethum, “Compressibility Moduli for Porous Materials Incorporating
Volume Fraction with Detailed Computations.” February 2006.
231 Jan Mandel, “Efficient Implementation of the Ensemble Kalman Filter.” May 2006.
... meteorological analysis, marine ecosystem modeling, etc.) (Evensen, 233 2003). Although the method relies on the Gaussian assumption, it is employed also in nonlinear 234 contexts (Mandel, 2006). 235 ...
Preprint
Full-text available
Researchers using crop models have been devising new roles for data and crop modeling based on the former’s increased availability and the new techniques developed for the latter. From the various available techniques, modeling may be tackled by data-driven methods or through a process-based approach. Process-based or mechanistic models may nonetheless take advantage of real-time observations through data assimilation. And while this approach has been widely used for field crops, this is not the case for crops grown in protected environments. We present a case study of data assimilation in a protected environment, capturing tomato growth data from different sources. We updated growth estimates of the Reduced State TOMGRO model, by assimilating observational data obtained through the continuous monitoring of plant mass and images captured by low-cost cameras, using the Unscented Kalman Filter and the Ensemble Kalman Filter. Since these techniques had not been used yet in the protected cultivation of tomatoes, it was necessary to develop the observation models as well, establishing the relationship between the observed variables and the ones estimated by the process-based model. The employed measurements, i.e., area of organs observed in pictures and plant-water mass, seemed suitable for tracking plant growth and for obtaining good approximations of the state variables estimated by the model. However, the quality of observations and of observation models was crucial for good performance of the assimilation techniques. As with other crops, it was not the case that assimilating one observation was useful for improving the value of others, including yield. We also observed that the assimilation performed better than calibrated models when there was a need to adjust the estimates to growth disturbances and that when filters lead to better yield estimates, continuous observations may not be required. There are then several steps and decisions that should be considered when bringing the idea from its application in field crops to protected environments and more studies are required to better determine the best approach.
... an equivalent representation from Mandel (2006) is used instead: ...
Article
Full-text available
A large fraction of costs in wild fisheries are fuel related, and while much of the costs are related to gear used and stock targeted, search for fishing grounds also contributes to fuel costs. Lack of knowledge on the spatial abundance of stocks during the fishing season is a limiting factor for fishing vessels when searching for suitable fishing grounds, and with better planning and routing, costs can be reduced. Strategic and tactical decision-making can be improved through operational decision support tools informed by real-time data and knowledge generated from research. In this article, we present a model-based estimation approach for predicting catch potential of ocean areas. An individual-based model of herring migrations is combined with an estimation approach known as Data Assimilation, which corrects model states using incoming data sources. The data used to correct the model are synthetic measurements generated from neural network output. Input to the neural network was vessel activity data of over 100 fishing vessels from 2015-2018, targeting mainly herring. The output is the predicted normalized density of herring in discrete grid cells. Model predictions are improved through assimilation of synthetic measurements with model states. Characterizing patterns from model output provides novel information on catch potential which can inform fishing activity.
... 1) Equations: The implementation of the EnKF follows the formulation presented in [2], [20]- [22] with a minor difference in the representation of the measurement error introduced by [23]. ...
... The P matrix is therefore of a size M ≪ N. For high-dimensional updates M ≫ N, it is shown in [30] how to reformulate P to obtain a linear complexity in M. ...
Article
Full-text available
Magnetic fields generated by normal or superconducting electromagnets are used to guide and focus particle beams in storage rings, synchrotron light sources, mass spectrometers, and beamlines for radiotherapy. The accurate determination of the magnetic field by measurement is critical for the prediction of the particle beam trajectory and hence the design of the accelerator complex. In this context, state-of-the-art numerical field computation makes use of boundary-element methods (BEM) to express the magnetic field. This enables the accurate computation of higher-order partial derivatives and local expansions of magnetic potentials used in efficient numerical codes for particle tracking. In this paper, we present an approach to infer the boundary data of an indirect BEM formulation from magnetic field measurements by ensemble Kálmán filtering. In this way, measurement uncertainties can be propagated to the boundary data, magnetic field and potentials, and to the beam related quantities derived from particle tracking. We provide results obtained from real measurement data of a curved dipole magnet using a Hall probe mapper system.
Article
Statistical learning additions to physically derived mathematical models are gaining traction in the literature. A recent approach has been to augment the underlying physics of the governing equations with data driven Bayesian statistical methodology. Coined statFEM, the method acknowledges a priori model misspecification, by embedding stochastic forcing within the governing equations. Upon receipt of additional data, the posterior distribution of the discretised finite element solution is updated using classical Bayesian filtering techniques. The resultant posterior jointly quantifies uncertainty associated with the ubiquitous problem of model misspecification and the data intended to represent the true process of interest. Despite this appeal, computational scalability is a challenge to statFEM's application to high-dimensional problems typically experienced in physical and industrial contexts. This article overcomes this hurdle by embedding a low-rank approximation of the underlying dense covariance matrix, obtained from the leading order modes of the full-rank alternative. Demonstrated on a series of reaction-diffusion problems of increasing dimension, using experimental and simulated data, the method reconstructs the sparsely observed data-generating processes with minimal loss of information, in both the posterior mean and variance, paving the way for further integration of physical and probabilistic approaches to complex systems.
Article
Full-text available
SUMMARY This paper describes a parallel preconditioned algorithm for solution of partial eigenvalue problems for large sparse symmetric matrices on massively parallel computers, taking advantage of advances in the scalable linear solvers, in particular in multigrid technology and in incomplete factorizations, developed under the Hypre project, at the Lawrence Livermore National Laboratory (LLNL), Center for Applied Scientic Computing. The algorithm implements a îmatrix freeî locally optimal block preconditioned conjugate gradient method (LOBPCG), suggested earlier by the rst author, to compute one or more of the smallest eigenvalues and the corresponding eigenvectors of a symmetric matrix. We discuss our Hypre specic implementation approach for a e xible parallel algorithm, and the capabilities of the developed software. We demonstrate parallel performance on a set of test problems, using Hypre algebraic multigrid and other preconditioners. The code is written in MPI based C-language and uses Hypre and LAPACK libraries. It has been included in Hypre starting with revision 1.8.0b, which is publicly available on the Internet at LLNL web site. Copyright c ∞ 2005 John Wiley & Sons, Ltd.
Article
Full-text available
Many geophysical problems are characterized by high-dimensional, nonlinear systems and pose difficult challenges for real-time data assimilation (updating) and forecasting. The present work builds on the ensemble Kalman filter (EnsKF), with the goal of producing ensemble filtering techniques applicable to non-Gaussian densities and high-dimensional systems. Three filtering algorithms, based on representing the prior density as a Gaussian mixture, are presented. The first, referred to as a mixture ensemble Kalman filter (XEnsF), models local covariance structures adaptively using nearest neighbors. The XEnsF is effective in a three-dimensional system, but the required ensemble grows rapidly with the dimension and, even in a 40-dimensional system, we find the XEnsF to be unstable and inferior to the EnsKF for all computationally feasible ensemble sizes. A second algorithm, the local-local ensemble filter (LLEnsF), combines localizations in physical as well as phase space, allowing the update step in high-dimensional systems to be decomposed into a sequence of lower-dimensional updates tractable by the XEnsF. Given the same prior forecasts in a 40-dimensional system, the LLEnsF update produces more accurate state estimates than the EnsKF if the forecast distributions are sufficiently non-Gaussian. Cycling the LLEnsF for long times, however, produces results inferior to the EnsKF because the LLEnsF ignores spatial continuity or smoothness between local state estimates. To address this weakness of the LLEnsF, we consider ways of enforcing spatial smoothness by conditioning the local updates on the prior estimates outside the localization in physical space. These considerations yield a third algorithm, which is a hybrid of the LLEnsF and the EnsKF. The hybrid uses information from the EnsKF to ensure spatial continuity of local updates and outperforms the EnsKF by 5.7% in RMS error in the 40-dimensional system.
Article
A theory for estimating the probability distribution of the state of a model given a set of observations exists. This nonlinear filtering theory unifies the data assimilation and ensemble generation problem that have been key foci of prediction and predictability research for numerical weather and ocean prediction applications. A new algorithm, referred to as an ensemble adjustment Kalman filter, and the more traditional implementation of the ensemble Kalman filter in which "perturbed observations" are used, are derived as Monte Carlo approximations to the nonlinear filter. Both ensemble Kalman filter methods produce assimilations with small ensemble mean errors while providing reasonable measures of uncertainty in the assimilated variables. The ensemble methods can assimilate observations with a nonlinear relation to model state variables and can also use observations to estimate the value of imprecisely known model parameters. These ensemble filter methods are shown to have significant advantages over four-dimensional variational assimilation in low-order models and scale easily to much larger applications. Heuristic modifications to the filtering algorithms allow them to be applied efficiently to very large models by sequentially processing observations and computing the impact of each observation on each state variable in an independent calculation. The ensemble adjustment Kalman filter is applied to a nondivergent barotropic model on the sphere to demonstrate the capabilities of the filters in models with state spaces that are much larger than the ensemble size. When observations are assimilated in the traditional ensemble Kalman filter, the resulting updated ensemble has a mean that is consistent with the value given by filtering theory, but only the expected value of the covariance of the updated ensemble is consistent with the theory. The ensemble adjustment Kalman filter computes a linear operator that is applied to the prior ensemble estimate of the state, resulting in an updated ensemble whose mean and also covariance are consistent with the theory. In the cases compared here, the ensemble adjustment Kalman filter performs significantly better than the traditional ensemble Kalman filter, apparently because noise introduced into the assimilated ensemble through perturbed observations in the traditional filter limits its relative performance. This superior performance may not occur for all problems and is expected to be most notable for small ensembles. Still, the results suggest that careful study of the capabilities of different varieties of ensemble Kalman filters is appropriate when exploring new applications.
Article
Ensemble data assimilation methods assimilate observations using state-space estimation methods and low-rank representations of forecast and analysis error covariances. A key element of such methods is the transformation of the forecast ensemble into an analysis ensemble with appropriate statistics. This transformation may be performed stochastically by treating observations as random variables, or deterministically by requiring that the updated analysis perturbations satisfy the Kalman filter analysis error covariance equation. Deterministic analysis ensemble updates are implementations of Kalman square root filters. The nonuniqueness of the deterministic transformation used in square root Kalman filters provides a framework to compare three recently proposed ensemble data assimilation methods.
Article
Traditional deformation parameters such as drained, unjacketed, and pseudo-bulk compressibilities are developed for a saturated porous media assuming they can be uniquely determined by measuring the liquid pressure, conflning pressure, and changes in volume. Physically however, the liquid volume fraction (porosity for a saturated porous medium) plays an important role, yet it is not directly measurable. In this paper the compressibilities are deflned in terms of the experiments used to evaluate them, and then mathematically related to the com- pressibilities of the liquid and solid phases and the volume fraction. The results can then be used to determine the efiects volume fraction and compressibilities of each phase have on deformation properties of the porous media. This theory is then shown to be a generalization of previous works and comparisons are made with previously derived relationships of Zimmerman, Biot, and Gassmann. This version contains details of computations in Appendix A.
Article
Abstract This paper is a follow-up to Dean and Russell[2] in which the formulation of our model of NAPL ∞ow in heterogeneous porous media is based on the traditional argument of developing a Fokker-Planck equation for a difiusion process and then modifying this equation to handle the behavior of the ∞uid particle at an interface between sands of difierent permeabilities. Since the capillary difiusivity can change signiflcantly across the interface, we call these changes jumps and incorporate them,into the stochastic difierential equation that describes the motion of the NAPL particle. Since this is a traditional approach that is modifled to incorporate non- traditional behavior such as jumps, we call it a bottom-up approach. The modifled stochastic difierential equation can then model pooling and channeling that occurs at such interfaces. In this paper, an attempt is made to accomplish the same result starting with a theory of stochastic processes that allows jumps. We start with semimartingale processes which are cµ adlµ ag. Since these processes allow jumps by deflnition, we call this approach the top-down approach.
Article
Ensemble data assimilation methods assimilate observations using state-space estimation methods and low- rank representations of forecast and analysis error covariances. A key element of such methods is the transfor- mation of the forecast ensemble into an analysis ensemble with appropriate statistics. This transformation may be performed stochastically by treating observations as random variables, or deterministically by requiring that the updated analysis perturbations satisfy the Kalman filter analysis error covariance equation. Deterministic analysis ensemble updates are implementations of Kalman square root filters. The nonuniqueness of the deter- ministic transformation used in square root Kalman filters provides a framework to compare three recently proposed ensemble data assimilation methods.
Article
Knowledge of the probability distribution of initial conditions is central to almost all practical studies of predictability and to improvements in stochastic prediction of the atmosphere. Traditionally, data assimilation for atmospheric predictability or prediction experiments has attempted to find a single ''best'' estimate of the initial state. Additional information about the initial condition probability distribution is then obtained primarily through heuristic techniques that attempt to generate representative perturbations around the best estimate. However, a classical theory for generating an estimate of the complete probability distribution of an initial state given a set of observations exists. This nonlinear filtering theory can be applied to unify the data assimilation and ensemble generation problem and to produce superior estimates of the probability distribution of the initial state of the atmosphere (or ocean) on regional or global scales. A Monte Carlo implementation of the fully nonlinear filter has been developed and applied to several low-order models. The method is able to produce assimilations with small ensemble mean errors while also providing random samples of the initial condition probability distribution. The Monte Carlo method can be applied in models that traditionally require the appli- cation of initialization techniques without any explicit initialization. Initial application to larger models is promising, but a number of challenges remain before the method can be extended to large realistic forecast models.
Article
Sign errors and typo's have been corrected. Abstract Coupling flow and deformation is of interest when modeling settling due to pump-ing water or oil, compaction due to construction, or, in other fields, modeling soft tissue, bones, or man-made porous materials. The modeling usually combines the Conservation of Mass equations for all phases, Darcy's law for flow, the Terzaghi's principal of effective stress, and a stress-strain relation. Here we discuss the assump-tions and thermodynamic framework underlying this formulation for single-phase flow and discuss the need to generalize the traditional model.