• Nie Znaleziono Wyników

Improving the operational forecasting system of the stratified flow in Osaka Bay using an ensemble Kalman filter–based steady state Kalman filter

N/A
N/A
Protected

Academic year: 2021

Share "Improving the operational forecasting system of the stratified flow in Osaka Bay using an ensemble Kalman filter–based steady state Kalman filter"

Copied!
19
0
0

Pełen tekst

(1)

Improving the operational forecasting system

of the stratified flow in Osaka Bay using

an ensemble Kalman filter–based

steady state Kalman filter

Ghada Y. H. El Serafy1and Arthur E. Mynett1,2,3

Received 8 August 2006; revised 16 October 2007; accepted 19 November 2007; published 20 June 2008.

[1] Numerical models of a water system are always based on assumptions and simplifications that may result in errors in the model’s predictions. Such errors can be reduced through the use of data assimilation and thus can significantly improve the success rate of the predictions and operational forecasts. The ensemble Kalman filter (EnKF) is a generic data assimilation method which is suited for highly nonlinear models. However, for three-dimensional operational systems such as in the case of Osaka Bay, Japan, a full EnKF would be computationally too demanding. In the present paper, a steady state Kalman filter (SSKF) simplification based on the correlation scales derived from the EnKF is proposed. This EnKF-based SSKF (EnSSKF) as presented in this paper is applied in combination with the three-dimensional Delft3D-FLOW system, modeling the stratified circulation system of Osaka Bay in Japan. The aim of the application of the EnSSKF is to improve the daily operational forecasts of salinity and current profiles for engineering activities within the basin. Salinity and velocity components were assimilated on an hourly basis for the period 13 –28 February 2002. The results of the filter performance and its forecasting ability are presented. The performance of the EnSSKF for improving the profiles of salinity and velocity components forecast during the first 24 h forecast is illustrated.

Citation: El Serafy, G. Y. H., and A. E. Mynett (2008), Improving the operational forecasting system of the stratified flow in Osaka Bay using an ensemble Kalman filter – based steady state Kalman filter, Water Resour. Res., 44, W06416,

doi:10.1029/2006WR005412. 1. Introduction

[2] The operational forecasting system of the stratified

Osaka Bay of Japan is used to predict the flow field and the salinity within the basin to assess the environmental impact of an ongoing construction within the bay [Tanaka et al., 2006]. These construction activities result in high concen-tration of suspended sediment outside the construction site that can be induced mainly by the flow field advected by relatively high velocity. To minimize the spreading of sedi-ments during the works, silt screens at different depths around the site are implanted; however, the turbidity around the construction should not exceed an environmental im-posed threshold of 2 mg/L. This operational forecasting system is further set up to predict the flow field to ensure those environmental regulations and to plan the operational activities in Osaka Bay. Measurements around the area of constructions where also collected for calibration and

val-idation purposes. The three-dimensional Delft3D-FLOW system, is used to model the stratified circulation system of Osaka Bay in Japan. The hydrodynamic model is being applied daily to provide 3-D current forecasts in a number of specified locations around the construction area. The objective of this research is to increase the accuracy of the 3-D current forecasts in the Osaka Bay through possible use of measurements.

[3] Hydrodynamic models often contain several sources

of uncertainty, which can occur at several stages during operation of the model. The governing equations may contain errors due to lack of knowledge about the complex physical processes and their interaction. Also, simplifica-tions often must be made to prevent high computation times. These simplifications will increase the model’s un-certainty. Uncertainties can also occur because of incorrect or incomplete input data of the model, such as boundary conditions, meteorological data, and bathymetry. The meas-urements on the other hand are often sampled with a low spatial and temporal resolution. To reduce those uncertain-ties in the model output and improve its predictions, data assimilation techniques such as Kalman filter techniques can be applied. Those techniques combine the model forecast with recent measurement data, using the informa-tion on the uncertainties in the model and the measurements to give a better estimate of the model output. An overview of the progress made in the field of ocean data assimilation,

1

Strategic Research and Development, WL Delft Hydraulics, Delft, Netherlands.

2

Environmental Hydroinformatics, Institute for Water Education, UN-ESCO-IHE, Delft, Netherlands.

3

Environmental Hydrodynamics, Delft University of Technology, Delft, Netherlands.

Copyright 2008 by the American Geophysical Union. 0043-1397/08/2006WR005412$09.00

W06416

WATER RESOURCES RESEARCH, VOL. 44, W06416, doi:10.1029/2006WR005412, 2008 Click

Here for

Full Article

(2)

singular evolutive interpolated Kalman (SEIK) filter [Pham et al., 1998]. A review of ensemble-based Kalman filters including the ensemble optimal interpolation is provided by Evensen [2003]. Since the physical processes at hand is highly nonlinear, the EnKF would be a sound choice for improving the flow field forecast. However, because of operational constraints, it is foreseen that though generic and easy to implement, the memory size and the computa-tion effort will be a burden for such a complicated stratified basin and under operational use. Simplifications have to be assumed. Reducing the computational cost using a simpler description of model dynamics can be done by using a coarser grid for the error covariance modeling in the numerical model [Cohn and Todling, 1996], or by approx-imating time consuming elements of the numerical model, such as employing cheaper numerical schemes, simpler turbulence closure schemes or assuming geostrophic bal-ance for the error covaribal-ance propagation [Dee, 1991]. However, the accuracy of the forecast is normally signifi-cantly deteriorating. Assuming a time invariant system, the steady state Kalman filter [SSKF] described by Morf et al. [1974] can be used to reduce the computational effort. Fukumori and Malanotte-Rizzoli [1995] derives such a steady gain from limiting theory solving the time invariant Riccati equation. Optimal interpolation (OI) has been also used by Kurapov et al. [2005], where the choice of the forecast error covariance is estimated on the basis of the ensemble statistics as described by Kurapov et al. [2002]. Also including geostatistical extensions in the structure of the uncertainties is described by Bertino et al. [2002].

[5] In this paper, and for the considerably large scale at

hand, of the stratified basin of Osaka Bay, an EnKF-based SSKF is proposed (EnSSKF). As applied by El Serafy et al. [2005, 2006, 2007], the structure of the uncertainties in the model and the typical correlation scales are obtained on the basis of calculations by the EnKF. The EnSSKF algorithm is here discussed in details and the choice of the correlation scales is also verified. Moreover, a full assessment of the improvements achieved in the daily operational forecasts of salinity and current profiles for environmental impact in this stratified basin of Osaka bay is here given.

[6] First, a description of the stratified three-dimensional

circulation system of Osaka bay forecasting system (OBFS) modeled by the three dimensional hydrodynamic Delft3D-FLOW system [Lesser et al., 2004] is given followed by the assumptions and techniques employed to define the corre-lation scales used in the data assimicorre-lation technique, viz. the EnKF-based steady state Kalman filter. The results from assimilating the salinity and velocity components are pre-sented for the period of 13 – 8 February 2002 using the

east by the Japan mainland, and to the west by the island Awaji. Osaka Bay is connected to the Seto Inland Sea by a wide and deep channel north of Awaji and a constriction with islands to the south. The primary driving force for the circulation is tide, which is mainly diurnal, with spring range in the order of 2 m. The monsoon-induced seasonal tilting of the ocean basin leads to slow variations in mean sea level in Osaka Bay, which are modeled by prescribing annual and semiannual ‘‘tidal’’ variations. In the northeast-ern part, five rivers discharge into Osaka Bay. Their discharge varies from very low to over 500 m3/s, leading to local salinity stratification in a three-dimensional circu-lation system. Salinity levels at the surface in that area vary from 25 to 32.5 psu (with more or less fully mixed conditions at the transition to the Seto Inland Sea). Varying wind and river discharges drive the local salinity variations. The Delft3D modeling system was used to model the area. A coarse grid overall 3-D model, with tidal open boundaries beyond the channels and outside Osaka Bay, was designed first. Figure 1 presents the model grid, channels, modeled islands, locations of open boundaries and rivers, etc. A detailed model for the most northeastern area is nested in this overall model. For this second model, a curvilinear orthogonal planar grid was designed, with a 45° clockwise rotated grid orientation (149  104 grid points with 10 equidistant vertical, so-called sigma, layers). The average cell size in the curvilinear fine grid varies between 20 to 40 m in the X direction and 30 to 50 m in the Y direction in the area of interest (i.e., around the construction area). Along the northwest open boundary of the nested model the current profile is prescribed. Along the southwest boundary, the water level is prescribed. Daily river dis-charges are provided by water level gauge recordings. A spatially uniform wind forcing is prescribed on an hourly basis. During the first part of the two month spin-up period for stratification, the wind variations are given at 24 h intervals. The model has been calibrated on boundary conditions and on tidal components. The calibration of the model setup, though a very essential step in the model evaluation, is not included in this paper. More details are given by Tanaka et al. [2006]. The stations (1, 2, 3, 4, and 5) shown in Figure 1 provide measurements of salinity and velocity components at 4 different vertical levels (1 m, 3 m and 6 m below the surface level, and 1 m from the bed level, respectively) for the period 13 – 28 February 2002. This set of field data is provided by Osaka Bay Regional Offshore Environmental Improvement Center [2002]. However, only stations 3 and 4 were used in the assimilation scheme, the other 3 stations were disregarded because of data nonavail-ability during the operational phase, and thus not used. Their data were used in this paper for verification of the

(3)

behavior of the proposed assimilation algorithm EnSSKF at unmeasured locations.

3. EnSSKF Algorithm

[8] The Kalman filter (KF) as introduced by Kalman

[1960] and Kalman and Bucy [1961] gives an optimal solution of the estimate with least variances (i.e., high accuracy). For nonlinear models, the sequential extended Kalman filter (EKF) is used which is fully described in [Jazwinsky, 1970]. However, the implementation of the EKF in a model requires knowledge of the detailed program description to derive a tangent linear version of the nonlin-ear model which leads to a model specific implementation. Another sophisticated data assimilation techniques based on KF for highly nonlinear models is the generic ensemble Kalman filter (EnKF) introduced by Evensen [1994b]. The EnKF is a sequential data assimilation method where the error statistics are predicted using Monte Carlo or ensemble integration [Evensen, 1994a]. The ensemble Kalman filter algorithm can easily be implemented for use with complex highly nonlinear models [Evensen, 1997] and is described as follows.

3.1. Ensemble Kalman Filter

[9] The nonlinear hydrodynamic model propagates the

system space state vector, xkin time. At initial time, tk, an

ensemble of size N is generated on the state vector. The ensemble is generated with a mean representing the initial condition of the state vector, ^xkjk1, and with a covariance

matrix that represents the uncertainty in the estimate of the initial condition, Pkjk1. The initial condition of the state

vector, ^xkjk1, is the estimate of the state vector at time k

conditioned on the measurements until time k 1. At every time step, tk, each ensemble member, i, with its state vector,

xkjk1i , forced by model errors, wki, is propagated in time

through the model. The model errors, wki, are randomly

drawn from a predefined distribution with zero mean and a covariance matrix, Qk. This covariance matrix represents

the structure of the uncertainties in the model (also addressed as model errors). These model errors can be regarded as the system noise. The estimate of the time update of the state vector can be calculated, at any time step, through the mean of the ensemble as

^ xkjk1¼ 1 N Xi¼N i¼1 xikjk1 ð1Þ

The error covariance matrix in the estimate of the time update of the state vector, Pkjk1, is calculated as a

covariance matrix of the ensemble Pkjk1¼ Lkjk1L0kjk1 where Lkjk1¼ 1 ffiffiffiffiffiffiffiffiffiffiffiffi N 1 p X N i¼1 xikjk1 ^xk1   ð2Þ

and L0kjk1 is the transpose of Lkjk1. Moreover, random

perturbations are added to the measurements in order not to violate the assumption of having the measurements as a random variable. An ensemble of size N of possible observations, yk

i

, is generated on the actual observations, using measurement errors, vk. The measurement errors are

also randomly generated from a predefined distribution with zero mean and covariance matrix, Rk, representing the

uncertainties in the measurements or measurement errors. The Kalman gain is then calculated using the measurement operator H that maps the state vector to measurement domain

Kk¼ Pkjk1H0kHkPkjk1H0kþ Rk 1

ð3Þ Furthermore, the updated state vector for every ensemble is then calculated as

xikjk¼ xi

kjk1þ Kk yik Hkxikjk1

 

ð4Þ Finally, using the results of equation (4), the estimate of the measurement update and the error covariance matrix are calculated as the mean of the ensemble, ^xkjk, and the

covariance of the ensemble, Pkjk. The advantage of the

ensemble Kalman filter is the feasibility of fast implementa-tion in complex high nonlinear models. One of the disadvantages of the EnKF is the computational effort and memory size as shown for one-dimensional hydrodynamic applications by El Serafy and Mynett [2004]. From this perspective, it is foreseen that the memory size and the computation effort will be a burden for such a complicated stratified basin and under operational use. Simplifications are thus necessary. Assuming a time invariant system, the steady state Kalman filter (SSKF) described by Morf et al. [1974] can be used to reduce the computational effort. 3.2. Steady State Kalman Filter

[10] The steady state assumes a time invariant system.

Thus a steady state covariance matrix and a steady state Kalman gain matrix are also time invariant as follows,

Pkjk1! P1;Pkjk! P1; and Kkjk1! K1 ð5Þ The above equations of equation (1) to equation (4) are then reduced to the following update equation

xkjk¼ xkjk1þ K1yk Hkxkjk1 ð6Þ Once measurements are available, the steady state Kalman gain matrix is used to update the system. The disadvantage of the steady state Kalman filter that effort should be done to verify the assumptions made on the Figure 1. Grid of the overall and nested Delft3D-FLOW

model for Osaka Bay.

(4)

Kalman gain structure which describes the spatial variability of the system.

3.3. EnKF-Based SSKF

[11] In this paper, a steady state Kalman gain matrix, K1,

based on the correlation scales of the covariance matrix, Pkjk1, calculated by the EnKF was used. In a

preopera-tional phase, the EnKF was used to calculate the covariance matrix during the time period of interest in which the measurements are available. The dominant correlation scales of the covariance matrix were extracted and used to construct the so-called steady state Kalman matrix. During the operational phase, this steady state Kalman gain matrix is used to update and improve the daily operational forecasts of salinity and current profiles for engineering activities in this stratified basin at acceptable computational effort.

[12] In section 3.3.1, the structure of the uncertainties as

assumed on the system noise used in the EnKF in the preoperational phase is given. Those assumptions are fur-ther verified in section 4.4. Some examples of the spatial distribution of the Kalman gain matrix used for the Osaka Bay are also given and discussed in section 3.3.2.

3.3.1. Role and Quantification of Uncertainties [13] A key aspect in the application of data assimilation is

the use of known or assumed uncertainties or errors in both measurements and process model. The uncertainties in the model and the correlation scales are often assumed on the basis of the experience with the model itself during the calibration and validation of the model setup. For the hydrodynamic model (Delft3D-FLOW), the uncertainties were assigned only to the water level, the salinity and the horizontal velocity components (i.e., independent variables in the filtering sense). All other variables such as vertical velocities, viscosity, and turbulence parameters are consid-ered dependent variables.

[14] Though the dimension of the Osaka bay is 19 km in

the X direction and 17 km in the Y direction, the branched area is almost half of the modeled domain. The area of interest lies around station 3 and station 4 (i.e., outside the branched area) and its dimension is around 9 km in the X direction and 8 km in the Y direction. The spatial correla-tion scales are roughly assumed within the dimension of the area of interest. No perturbations were induced on the boundaries of the modeled area since this might accumulate in time producing unrealistic oscillations that might affect the model stability. The deterministic simulation of the

detailed model for the full period of simulation (i.e., February 2002), was a guiding factor for the reasoning behind some of the correlation scales assumptions. For example, since the dominant tidal cycle is diurnal, it is expected to have a strong spatial horizontal correlation of the uncertainty in the salinity and in the water level. For the salinity and water level, the correlation length in the horizontal direction is roughly assumed to be 900 m and 800 m in the X direction and Y direction, respectively. Moreover, since the salinity profile during this period can change from stratified to well-mixed salinity conditions, choosing a specific vertical correlation for the salinity profile can be misleading. Thus no correlation was assumed for the uncertainties induced on the salinity profiles.

[15] The deterministic model results for the currents show

a high variability of the surface layer current due to wind fluctuations. This suggested weak spatial horizontal corre-lation for the uncertainties in the velocity components. For the currents, the correlation length used is 600 m and 500 m in X and Y directions, respectively. In the vertical, the wind effect is strongly damped in the water column. In the near-bed layer, tidal variations dominate but the current can be opposite to the surface current implying a vertical circula-tion. Because of the pronounced variation in the currents, it was assumed that there is no correlation in the vertical for the uncertainties in currents. Those assumptions lead to the definition of the uncertainties used to generate the induced noise on the ensemble for the EnKF. The noise generated was a normally distributed random white noise. The uncer-tainty induced on all variables was assumed uncorrelated in the vertical. An example of the normal distribution of the noise used on the salinity and the velocity components at station 3 is shown in Figure 2.

[16] It has to be emphasized that those assumptions are

rough assumptions but are based on experience with the typical model errors during the calibration phase and the dominant physics of the modeled phenomena as here explained. Any other assumptions would have an effect on the results here presented. For example, using unrealistic weaker spatial correlation might give a local update effect. On the other hand, introducing unrealistic stronger spatial correlation can lead to unrealistic results through the updat-ing step of the filter. The assumptions on the correlation scales here stated is further verified in section 4.4. Different correlation scales are assumed and used to define the Figure 2. Assumed distribution of the uncertainties in the salinity and the velocity components at

(5)

statistical structure of uncertainties. The effect of the use of different correlation scales than those here defined on the performance of the EnSSKF filter at unmeasured location are presented.

3.3.2. Defining the Correlation Scales for the OBFS [17] The EnKF was used to provide a rough estimate on

the correlation scales and thus the steady state Kalman gain used in Osaka Bay. An ensemble of 30 members was created by perturbing the deterministic solution every hour with the noise terms generated with the correlation scales defined in section 3.3.1, having a standard deviations equal to 0.5 psu for salinity and 1 cm/s for each of the velocity components. Each ensemble member (i.e., each member is representing the system space state vector at a given time) was propagated through the model in time. At every hour, the covariance matrices of equation (2) and the Kalman gain matrix as defined in equation (3) were calculated. At the end of the simulation period (28 February), those covariances (or Kalman gains) were averaged in time, to produce the steady state Kalman gain for the OBFS used in equation (6). By averaging in time over a period of 15 d (i.e., 360 h), the time-dependent correlation variations are smeared out, leav-ing the dominant correlations in the system. Those will be still present in the steady state Kalman gain matrix. For model stability reasons, further modifications were made to the averaged Kalman gain matrix. Though in the model, the salinity values in the middle of the river branches Ohkawa and Yamato were taken as a constant (i.e., of fresh water), very high values in the Kalman gain for the salinity field were observed in those branches. This indicates that the grid in those branches is unnecessarily dense introducing

unphysical influence of the measurements at station 3 and station 4 at the middle of the river branches with a constant salinity value of fresh water. The Kalman gain in the branches of the OBFS is set to zero. The Kalman gain represents the influence of each measurement on the domain at hand with normally a maximum influence at the mea-sured location. To get an indication of the spatial effect or here addressed as the correlation distribution in the Kalman gain used in updating the state through equation (6), each column was normalized to the gain value of the corresponding location. Examples of the correlation distri-bution at station 3 in the Kalman gain just described are illustrated in Figures 3, 4, 5, 6, 7, and 8 for horizontal correlation and for vertical correlation, respectively. For station 4, similar correlations were found.

[18] Figure 3 shows that the positive horizontal

correla-tion distribucorrela-tion of the salinity is more spread at the middle layers (i.e., from 3 m below the surface level to 6 m below) than at the top layer and the bottom layer. The negative horizontal correlation distribution is higher also at those layers. The mixing process is more active at the middle layers allowing for the propagation of the uncertainty. There is also more spread of the correlation in the X direction than in the Y direction. This might be due to the dominant velocity direction. The correlation profile for the salinity shown in Figure 6 is also smooth and seems to indicate that the correlation in the vertical is not high but can extend to 2 layers below the considered layer.

[19] Figures 4 and 5 show the spread of the horizontal

correlation distribution for the velocity components is less than that of the salinity, which is to be expected as explained Figure 3. Horizontal correlation distribution in the Kalman gain matrix for the salinity at measured

layers (i.e., 1, 3, and 6 m from the surface and 1 m from the bottom) at station 3.

(6)

Figure 4. Horizontal correlation distribution in the Kalman gain matrix for the velocity component in the U direction at measured layers (i.e., 1, 3, and 6 m from the surface and 1 m from the bottom) at station 3.

Figure 5. Horizontal correlation distribution in the Kalman gain matrix for the velocity component in the V direction at measured layers (i.e., 1, 3, and 6 m from the surface and 1 m from the bottom) at station 3.

(7)

Figure 7. Vertical correlation scales in the Kalman gain matrix for the velocity component in the U direction at the measured layers (i.e., 1, 3, and 6 m from the surface and 1 m from the bottom) at station 3. Figure 6. Vertical correlation scales in the Kalman gain matrix for the salinity at the measured layers (i.e., 1, 3, and 6 m from the surface and 1 m from the bottom) at station 3.

(8)

in section 3.3.1. The radius of the high positive correlation (higher than 0.5) around station 3 at all layers is larger for the salinity than that of the velocity. In other words the area around station 3 where correlation (positive or negative) is encountered is more extended for the salinity than that for the velocity. It is also less smooth because of the variation in the velocity magnitude and direction. Strong correlation was also encountered at the boundary which suggests that the boundaries are too close to the area of interest which can have a negative effect on the performance of the filter. Those correlations at the boundaries were smoothed out for stability reasons. It has to be mentioned here, that though the measurements were given in the north and east direc-tions, the Delft3D model variables (i.e., velocity compo-nents) are defined in the direction of the grid (i.e., U and V directions) as shown in Figure 2 (right plot). The Kalman gain matrix has to reflect the effect of the measurements on the variables used by the model and updated. Thus the correlation distributions plotted here for the velocity com-ponents are representing the correlation distributions of the velocity components in U and V directions rather than the north and east directions.

[20] The correlation profile shown in Figures 7 and 8 for

the velocity components suggests that there is a small or no correlation in the vertical. The figures of the correlation profiles for 1 m (first layer of the model), 3 m (second layer of the model) and 6 m (fourth layer of the model) from the surface show that correlation between the level at hand and only one adjacent layer is approximately 0.1 and no correlation further up or down is encountered. For the bottom layer (layer 10 of the model), however, Figures 7 and 8 show significant correlation between the bottom layer

and the three above adjacent layers. This might be due to the stratification of the water column (i.e., presence of saline water at the bottom) and the vertical diffusion used in Delft3D-FLOW at the bottom.

[21] There is also a cross correlation encountered between

the salinity and the velocity components. However, because of the short period of applications (15 d), those were not so obvious to be able to identify a constant behavior in time, thus it was not taken into account. If the time length of the data was longer including longer periods of stratification and/or longer periods of well-mixed behavior, it would have been easier to identify a typical behavior of the cross correlation between salinity and velocity components. Moreover, since the timescale of the variations in the horizon-tal velocity components in both directions is very short, it was also expected that the cross correlation between them would not be well identified through the averaging process. The correlation distribution shown in Figures 7 and 8 were used in the steady state Kalman gain matrix for the EnSSKF.

4. Application of the EnSSKF to the OBFS

[22] In this section, the application of the EnKF-based

steady state Kalman filter (EnSSKF) as the data assimilation technique applied to OBFS is described. The structure of the uncertainties in the model and the typical correlation scales were assumed on the basis of those calculated by the EnKF, see sections 3.3.1 and 3.3.2 and will be here verified in section 4.4. The aim of the application of SSKF is to improve the daily operational forecasts of salinity and current profiles for engineering activities in this stratified basin. The calibra-tion is shown for the case of assimilacalibra-tion, on an hourly basis, Figure 8. Vertical correlation scales in the Kalman gain matrix for the velocity component in the V

(9)

of salinity and velocity components for the period from 13 – 28 February 2002. The results of the filter and its forecasting ability are presented. The performance of the EnSSKF for improving the salinity and velocity components forecast during the first 24 h forecast is illustrated.

4.1. Reduction in Model Error due to Assimilation [23] At every grid point of the nested model of the Osaka

Bay, the model calculates the water level, the salinity, and the velocity components in the two directions (north

direc-tion and east direcdirec-tion). As previously mendirec-tioned, only stations 3 and 4 shown in Figure 1 are used in the assimilation. The measurements of salinity and velocity components at 4 different vertical levels (1 m, 3 m and 6 m below the surface level, and 1 m from the bed level, respectively.) at those two stations are available for the period 13 – 28 February 2002. The measurements were assimilated in the nearest grid cell to the station location which is very well accepted regarding the small grid cell size used. In the vertical, since the model uses sigma layers, Figure 9. Salinity measurements, the salinity predicted by the model without assimilation, and the

full-period assimilation hindcast salinity at station 3 at a vertical level 1 m from surface water level.

Figure 10. Measured north velocity component against that predicted by the model without assimilation and the full-period assimilation hindcast at station 3 at a vertical level 1 m from surface water level.

(10)

the depth of each layer is not constant in time and is dependent on the water depth. For simplicity reasons, the measurements at different vertical layers were also assimi-lated in the corresponding layer of the model, thus no horizontal or vertical interpolation is used in the assimila-tion scheme. The measurements are available on an hourly basis, while the model runs with a time step of one minute. Those measurements were assimilated in a hindcast run. The results of the full-period EnSSKF hindcast, the meas-urements used and the model run without assimilation are shown in Figures 9, 10, and 11 for the salinity, the north velocity and the east velocity components, respectively, at the vertical level 1 m from surface level at station 3. From Figures 9, 10, and 11, it is clear that the hindcast including

full-period assimilation of recent measurement data follows the measurements better than the model without assimila-tion. The results at other layers and those of station 4 show the same behavior. As a measure of goodness of fit (GoF), and to quantify the reduction in the error due to assimila-tion, the difference between the model and the measure-ments normalized to the measuremeasure-ments and averaged over the period of measurements,emodel, is calculated as follows:

emodel¼ 1 K XK k¼1 Xmodelð Þ  Xtk measurementð Þtk Xmeasurementð Þtk         ð7Þ

where X is one of the variables (i.e., salinity, north velocity component, and/or east velocity component) and K is the Figure 11. Measured east velocity component against that predicted by the model without assimilation

and the full-period assimilation hindcast at station 3 at a vertical level 1 m from surface water level.

Table 1. Time-Averaged Percentage Erroreand Percentage Reduction in the Model Errorg for Stations 3 and 4 at All Four Vertical Layers

1 m From Surface Level 3 m From Surface Level 6 m From Surface Level 1 m Above Bed Level Station 3 Station 4 Station 3 Station 4 Station 3 Station 4 Station 3 Station 4

ein Salinity

Model without assimilation 4.41 3.51 2.01 1.44 1.29 0.63 0.52 0.68

EnSSKF hindcast 1.08 0.97 1.42 0.49 0.90 0.25 0.09 0.10

gEnSSKF hindcast 75.5 72.4 29.4 66.0 30.2 60.3 82.7 85.3

ein North Velocity Component (for Velocities > 5 cm/s)

Model without assimilation 97.4 59.5 80.5 88.8 75.2 101.6 49.7 84.1

EnSSKF hindcast 26.0 23.8 62.6 26.4 49.7 26.6 16.6 32.8

gEnSSKF hindcast 73.3 60.0 22.2 70.3 33.9 73.8 66.6 61.0

ein East Velocity Component (for Velocities > 5 cm/s)

Model without assimilation 122.7 125.8 66.8 61.0 56.5 58.4 43.5 34.4

EnSSKF hindcast 34.8 44.9 50.8 23.3 21.4 28.7 17.3 17.1

(11)

total number of time steps, tk, where measurements are

available (i.e., 361 time steps in the 15 d starting from 13 February at 0000 UT and ending at 28 February at 0000 UT included). However, for measured velocity components less than 5 cm/s, the normalized error as defined in equation (7) can give the wrong information on the accuracy (i.e.,

division by small number or zero). Moreover, for opera-tional use, the main interest of this research was in improving the prediction of the velocities over 5 cm/s (high concentration of suspended sediment outside the construction site can be induced mainly by the advection with relatively high velocity), thus smaller velocities were Figure 12. Salinity measurements against that predicted by the model without assimilation, the hindcast

with assimilation, and the 24 h forecasted salinity at station 3 at a vertical level 1 m from surface water level.

Figure 13. Measured north velocity component against that predicted by the model without assimilation, the hindcast with assimilation, and the 24 h forecast at station 3 at a vertical level 1 m from surface water level.

(12)

disregarded from the calculations in those cases. Similarly, for the hindcast, the percentage error, ehindcast, was

calculated. The percentage reduction in error is then defined as ghindcast = (emodel ehindcast)/emodel%. The errors in the

salinity, the north and the east velocity component together with reduction percentages, g, are presented in Table 1. These are shown for the two measured stations (i.e., stations 3 and 4) at the four measured layers. The overall reduction in the error varies between 29.4% – 85.3% for the salinity, 22.2% – 73.8% for the north velocity component and 24.0% – 71.6% for the east velocity component, respectively. The same percentages were also calculated for the absolute errors (not shown here) leading to very similar results. 4.2. Improvement of the Forecasting Capability of the Model

[24] The model is driven by tide, wind and fresh water

river discharge. The model is being applied daily to provide 3-D current forecasts in a number of specified locations. These operational forecasts are used for planning operation-al activities in Osaka Bay. The main objective was to increase the accuracy of the forecast using the EnSSKF. This predictive capability of the EnSSKF was compared to the results of the model without assimilation. Similar to situations in operational systems, it is assumed that the measurement data are available until a specific day and time (here, 19 February, 0000 UT) and a 24 h forecast is required (here, from 19 February, 0000 UT, until 20 February, 0000 UT). During the past period, viz. up to 19 February 0000 UT, the EnSSKF was provided with the set of measurements every hour from the beginning of the simu-lation time (13 February 2002), until the end of the assimilation period (19 February). Obviously, no assimila-tion of data is done beyond that date. The state is then predicted in a forecast mode until the 20 February, 0000 UT.

The results in the forecast period are addressed here as EnSSKF forecast. Since the model is driven by a diurnal tide, the main improvement of the forecasted state is expected within the first 12 h and its effect may rapidly die out with time. However, since the stratification in the salinity extends over a longer timescale, the improvement in the forecasted state can last longer. The EnSSKF forecast, the model without assimilation, the EnSSKF hindcast and the measurements are shown in Figures 12, 13, and 14 for the salinity, the north velocity and the east velocity compo-nents, respectively, at the vertical level 1 m from surface water for the period (16 – 20 February 2002), all for station 3. For station 4, similar results were also encountered. From Figures 12, 13, and 14, it is seen that forecast improvements are realized within the first 24 h for the salinity and 21 h for the north velocity. For the east velocity, this improvement is only within the first 6 – 10 h. This can be explained as follows: The boundary condition of the nested model is provided from the result of the overall coarse model. The effect of this on the east velocity component during this period (18 – 20 February) is dominant thus reducing the effect of the update on the east velocity faster than on the north velocity. Also from Figures 12, 13, and 14, it is observed that the improvement in the forecast of the salinity within the 24 h forecast is stronger than that of the velocity components. The results at other layers and at station 4 resemble those results, except for the bottom layer, which is to be expected. The bottom layer is always a bounded layer. Through the model, the short term temporal oscillations are being damped out in the bottom layer using the turbulence closure model, forcing the updated salinity to relax quickly to the model and according to the fluxes; the salinity might show a different behavior from the model without assimilation.

Figure 14. Measured east velocity component against that predicted by the model without assimilation, the hindcast with assimilation, and the 24 h forecast at station 3 at a vertical level 1 m from surface water level.

(13)

[25] The improvement in the forecast capability of the

model for this particular 24 h case is shown against the forecast duration for the salinity, the north velocity compo-nent and the east velocity compocompo-nents in Figures 15, 16, and 17. Figures 15 – 17 confirm that forecast improvements are realized within the first 24 h forecast for the salinity in Figure 15, and 21 h for the north velocity in Figure 16. For the east velocity, Figure 17, this improvement is only within the first 6 – 10 h. We note that these results on improvement are valid for this specific case of 24 h forecast. The results for another 24 h period may be different, that is, better or worse. In general, the improvements for a large number of 24 h forecasts need to be considered, to obtain the infor-mation on forecast improvement over the 24 h forecast window in a statistical sense, which is a robust quantitative measure for improvement. This is illustrated only for a 6 h forecast window in section 4.5. The present case serves

mainly as an illustration of the improvement and the method for a quantitative assessment of these improvements. 4.3. Performance Assessment at Unmeasured Locations

[26] Besides the measurements of stations 3 and 4, which

were used to show the improvement of the forecast when using data assimilation (section 4.2), measurement data for the same verification period 13 – 28 February 2002 are available for stations 1, 2 and 5. The improvements in those stations are due to the correlation scales used in the covariance matrix and given in section 3.3.2 and due to the propagation of the updated fields at the measured station (i.e., station 3 and 4) to the position of the ‘‘unmeasured’’ stations (i.e., station 1, 2 and 5) through the dynamics of the model. The available measurements at those locations can therefore be used for the assessment of the EnSSKF performance and for verification of the prescribed

correla-Figure 16. Improvement in the north velocity forecast capability of the model against the forecast duration for 19 February 2002, 0000 – 2400 UT.

Figure 15. Improvement in the salinity forecast capability of the model against the forecast duration for 19 February 2002, 0000 – 2400 UT.

(14)

tion scales. The time-averaged percentage reduction in the model error as defined in section 4.1 was also calculated for the unmeasured stations at the four different vertical levels and is presented in Table 2. Since the velocities at those stations were small most of the time, the threshold on the velocity was lowered to 1 cm/s. Improvement in the model error due to assimilation can be recognized by a positive percentage reduction, (emodel>ehindcast), while deterioration

is expressed as a negative percentage reduction (emodel <

ehindcast).

[27] Table 2 shows that improvements are also

encoun-tered at these locations for the salinity as well as for the velocity components. As expected, the improvements at unmeasured stations (away from the measured stations) are less than the improvements at the measured locations. At some locations, a slight deterioration occurs. At station 1 at the top layer (1 m from the surface level) and at station 2 (3 m from surface level) slight deteriorations of the updates in the salinity occur. This is due to the assumed correlation being a time-averaged value of the expected correlation scales, which may vary in time, thus resulting in slight deviations from the measurements. Also at station 2 (3 and 6 m from surface water), and station 5 for the bottom layer, a slight deterioration in the east velocity component is shown. Considerable deterioration is observed at station 5 (6 m from the surface) in the salinity and at station 2 at the bottom layer for the north velocity component.

[28] At all stations (6 m from the surface), the general

pattern of the salinity measurements is being followed, but the variability in the measurements is not being captured. Since the model without assimilation performs reasonably well at that layer and the EnSSKF has no knowledge of the variability of the data at unmeasured stations, the percentage reduction can become very small as for station 1 (0.68%) or reasonable as in station 2 (13.89%) or even negative as for station 5 (22.55%). For station 5, this high negative reduction was due to the fact that its correlation to the measured stations (station 3 and 4) in the Kalman gain is high (i.e., significant influence from the measured stations). Thus, the filter is correcting the salinity at station 5 with the variability of the data at stations 3 and 4 which is not the same in time.

[29] At station 1, the percentage reduction in the salinity

error is low, but in the velocity error is high. This can be explained by observing the horizontal distribution of the

correlation for the salinity and the velocity components in Figures 3 – 5. A very small correlation for the salinity distribution is encountered between the measured stations (station 3 and 4) and the unmeasured station 1 (i.e., varying between 0.05 and 0.08 for different levels), while for the velocity components it is considerably higher (i.e., can reach up to ±0.2 for the V velocity and ±0.12 for the U velocity).

[30] At station 2 (1 m above bed level), a substantial

deterioration of the north velocity component is encoun-tered (19.92%) against a considerable improvement in the salinity profile (40.16%) and a slight improvement in the east velocity component (8.13%). Since the correlation between the salinity and the velocity components is not taken into consideration, the gain (i.e., the improvement) in the salinity is not reflected on the velocity components of the unmeasured location resulting in slight update or update in the wrong direction at some periods.

[31] From the above it is concluded that the correlation

scales used are reasonable but can be further enhanced by including the correlation between the velocity components and the salinity. However, since the period of the available data was too short and the timescale of the variations in the horizontal velocity is also short, further enhancement on the correlation scales used was not feasible. Similarly, the results indicate that the steady state assumption of the length scales of correlation and therefore the covariance matrix and the Kalman gain matrix has its limitation, in particular for station 1 close to the boundary.

4.4. Effect of the Assumed Correlation Scales on the Performance at Unmeasured Stations

[32] The effect of the assumptions on the correlation scales

presented in section 3.3.2, is here verified. Though the assumptions are rough assumptions, any significantly dif-ferent assumptions would have an effect on the results presented. The improvement in the model results at unmea-sured stations (i.e., stations 1, 2 and 5) during the assimi-lation of the measurements at station 3 and station 4, are mainly due to the correlation scales used in the covariance matrix given in section 3.3.1. and due to the propagation of the updated fields at the measured station to the position of the ‘‘unmeasured’’ stations through the dynamics of the model. To ensure the validity of the assumptions on the correlation scales given in section 3.3.1. (i.e., the correlation Figure 17. Improvement in the east velocity forecast capability of the model against the forecast

(15)

length used is 900 m and 800 m in X and Y directions for the salinity and 600 and 500 m for the velocity components X and Y directions), different correlation lengths are as-sumed and used to define the statistical structure of uncer-tainties. Smaller correlation length is used in (case of short correlation) of 250 m and 200 m for the salinity and 150 and 125 m for the velocity in X and Y while in (case of long correlation) bigger correlation length are used of 3600 m and 3200 m for the salinity and 2400 and 2000 m for the velocity components. The model results (i.e., EnSSKF hindcast) for those two cases were compared to that of the EnSSKF hindcast of station 5. The results are shown in Figure 18 for station 5 for the salinity at 1 m below surface level. For the sake of comparison, the EnSSKF hindcast of

station 4 which is the nearest measured station to station 5 is also included in Figure 18. Similar results were found for the other levels and for other stations, thus are not here shown. For clarity, only 2 d are depicted within Figure 18. From Figure 18, it is shown that in case of short correlation scales, there is not that much influence on the results at station 5. The EnSSKF hindcast of station 5 resembles the model with no assimilation of station 5. With long correla-tion scales, the EnSSKF hindcast of stacorrela-tion 5 resembles the variation of the hindcast of station 4 (i.e., the nearest measured station). The results (i.e., EnSSKF hindcast at station 5) using correlation scales as in section 3.3.2 are closer to the ‘‘unassimilated’’ measurements more than the model without assimilation. From that it is concluded that Table 2. Time-Averaged Percentage Reduction in the Model Errorg for Stations 1, 2, and 5 at All Four Vertical

Layers

Percentage Reduction in Error in Model 1 m From Surface Level 3 m From Surface Level 6 m From Surface Level 1 m Above Bed Level g in Salinity Station 1 6.38 9.65 0.68 8.18 Station 2 1.71 5.80 13.89 40.16 Station 5 8.81 17.58 22.55 18.44

g in North Velocity Component (for Velocities > 1 cm/s)

Station 1 18.92 19.37 6.06 19.70

Station 2 14.38 14.11 7.83 19.92

Station 5 18.57 15.54 29.98 32.67

g in East Velocity Component (for Velocities > 1 cm/s)

Station 1 18.92 19.37 6.06 19.70

Station 2 8.50 4.07 7.24 8.13

Station 5 3.02 7.48 2.01 0.20

Figure 18. Salinity EnSSKF hindcast at station 5 with different correlation scales and the model without assimilation together with the measurements and the EnSSKF hindcast at station 4 for pattern comparison.

(16)

using unrealistic weaker spatial correlation might give a local update effect The radius of influence of the update from the measurements would be very small thus updating only at the measurement location. On the other hand, introducing unre-alistic stronger spatial correlation can lead to unreunre-alistic results through the updating step of the filter. The radius of influence of the update from the measurements would be very large thus close measurements might give contradictory effect on the update which will lead to unacceptable results. It is shown here that the correlation scales used in this research though roughly assumed are realistic assumptions. 4.5. Assessment of the Overall Accuracy of the 6 h Forecasting Capability of the Model

[33] Preliminary tests in section 4.2 showed that the

fore-cast improvement by the model with EnSSKF was effective only within 24 h at most. The need of controlling the turbidity produced as the results of construction activities and predicting the direction in which this turbidity may travel causing undesired pollution on coastal environment is

the main goal of this research, hence a 6 h forecast is proposed in which the accuracy of the forecast is assessed. To be able to define the 6 h forecast accuracy for the whole period, the latest hindcast at every hour is used for the following 6 h forecast. The latest hindcast is the hindcast calculated using the data from the beginning of the simu-lation until the time proposed for the forecast. Three parameters were defined to evaluate the improvement of the forecast and the duration of reasonable forecast ability: (1) the hit rate of current direction, (2) the hit rate of the current magnitude and (3) that for the salinity.

[34] The hit rate is defined as the rate of occurrence of a

‘‘successful’’ forecast (i.e., percentage of the number of successful occurrences). The successful forecast is defined per variable. When the absolute value of a deviation angle between the forecasted and observed current directions is less than 45°, the forecast is judged as being ‘‘successful.’’ The successful forecast for the current magnitude is defined such that the absolute value of the error in the velocity magnitude should not exceed 0.05 m/s (i.e., 5 cm/s). The

3 h 72 56 60 70 65 6 h 69 49 55 65 60 Average Model 64 47 43 59 53 EnSSKF 94 82 99 100 94 3 h 71 58 73 75 69 6 h 67 56 68 72 66

Table 4. Magnitude Hit Rate for Model Without Assimilation and EnSSKF Hindcast, 3 h Forecast, and 6 h Forecast for Measured Locations Station 3 and Station 4

1 m From Surface Water 3 m From Surface Water 6 m From Surface Water 1 m Above Bottom Level Vertical Average Station 3 Model 34 76 82 85 69 EnSSKF 81 72 90 100 86 3 h 61 91 91 100 86 6 h 52 85 85 100 81 Station 4 Model 32 56 75 76 60 EnSSKF 81 90 97 98 92 3 h 39 93 97 99 82 6 h 38 91 96 99 81 Average Model 33 66 79 81 65 EnSSKF 81 81 94 99 89 3 h 50 92 94 100 84 6 h 45 88 91 100 81

(17)

judgment was done only when the absolute value of the observational velocity is more than 5 cm/s, because high concentration of suspended sediment outside the construc-tion site can be induced mainly by the advecconstruc-tion with relatively large velocity. For the salinity, since the variation of 0.5 psu does not have that much influence on the velocity distribution, the forecast is assumed successful when the absolute error in the salinity is less than 0.5 psu. In order to evaluate the accuracy of the predicted variables quantita-tively, the rate of occurrence of ‘‘successful’’ forecast, here addressed as the hit rate, is calculated for the whole period of 13 – 28 February 2002 for each layer, for each station and for each variable separately.

[35] The hit rates for the current direction, the current

magnitude, and the salinity as here defined, are shown in Tables 3, 4, and 5, respectively, for the model with no assimilation, EnSSKF hindcast, 3 h forecast and 6 h forecast, respectively. Those are presented for each layer of the measured stations (stations 3 and 4).

[36] From Tables 3, 4, and 5, it can be inferred that the hit

rates are improved when using EnSSKF at every depth. The 3 h forecast and the 6 h forecast show higher hit rates than those of the model with no assimilation. In general, the hit rate of EnSSKF hindcast decreases in time as expected (i.e., the hit rate of 6 h forecast is lower than that of the 3 h forecast which in turn is lower than that of the EnSSKF hindcast). The maximum hit rates are observed at the middle and lower layers because of the effect of vertical mixing in the well mixed part of the stratified zone. Those layers are almost always below the salinity isolines. At the surface layer, the hit rates of the EnSSKF hindcast decreases rapidly in the forecasting process. Since the current in the surface layer is dominated by wind forcing and by the depth of mixing layer, it is determined through complicated hydro-dynamic processes. The accuracies of wind and river dis-charge data directly affect the forecast accuracy for the surface current. On average, the direction hit rate of the model with no assimilation of 53% is lower than the magni-tude and salinity hit rates (65% and 64% respectively.). This increases to 94%, 89% and 90% for the EnSSKF hindcast. For the 3 h forecast values of 69%, 84%, and 78% were

observed. For the 6 h forecast, hit rates of 66%, 84%, and 75% were calculated. Those hit rates are considered accept-able for operationally accuracy. However, the rapid decrease of the direction hit rates in the forecasting mode, unlike the magnitude and salinity hit rates, seems to indicate that the model’s predictive capability for velocity components (i.e., north and east components) is not as accurate as the overall prediction capability for salinity and current magnitude.

5. Summary and Conclusions

[37] In this paper, the deterministic Delft3D-FLOW

hy-drodynamic model is extended with an ensemble Kalman filter (EnKF)-based steady state Kalman filter technique (EnSSKF) that enables assimilation of recent observational data, thus improving the forecasting capability of the Osaka Bay Forecasting System. A steady state Kalman matrix based on the covariances calculated by the EnKF was defined. The correlation distribution present in the Kalman gain matrix reflected the dynamics of the Osaka Bay Forecasting System. It suggested that the boundaries are too close to the area of interest. It also reflected that the computational grid in the branches is unnecessarily dense.

[38] The aim of the application of EnSSKF is to improve

the daily operational forecasts of salinity and current pro-files in this stratified basin. Salinity and velocity compo-nents were assimilated for the period 13 – 28 February 2002 on an hourly basis. The results show that the EnSSKF hindcast follows the measurements better than the model without assimilation. To quantify the reduction in the error due to assimilation, the percentage reduction in error (i.e., a measure of goodness of fit) was defined on the basis of the relative differences between observations and model pre-dictions. The overall reduction in the error varies between 29.4% – 85.3% for the salinity, 22.2% – 73.8% for the north velocity component and 24.0% – 71.6% for the east velocity component.

[39] The results also show the forecast improvement due

to assimilation. The results of EnSSKF forecast are better than those of the model without data assimilation, with the update effect disappearing gradually for larger forecast times. The forecast improvement occurs within the first Table 5. Salinity Hit Rate for Model Without Assimilation and EnSSKF Hindcast, 3 h Forecast, and 6 h

Forecast for Measured Locations Station 3 and Station 4

1 m From Surface Water 3 m From Surface Water 6 m From Surface Water 1 m Above Bottom Level Vertical Average Station 3 Model 32 45 66 96 60 EnSSKF 82 67 84 100 83 3 h 52 61 78 100 73 6 h 43 56 79 100 70 Station 4 Model 21 64 94 91 68 EnSSKF 84 98 100 100 96 3 h 47 86 97 100 83 6 h 41 83 95 99 80 Average Model 27 55 80 94 64 EnSSKF 83 83 92 100 90 3 h 50 74 88 100 78 6 h 42 70 87 100 75

(18)

these ‘‘unmeasured’’ stations (i.e., stations 1, 2 and 5) through the dynamics of the Delft3D-FLOW model. As expected, the improvements at unmeasured stations are smaller and sometimes even negative compared with the improvements at the measured location. This is due to the assumed correlation that is a time-averaged value of the expected correlation scales, which may vary in time, thus resulting in deviation from the measurements. However, in general, results showed that the correlation scales used are reasonable but can be further enhanced by including the correlation between the velocity components and the salin-ity. Since the period of the available data was too short, further enhancement on the correlation was not feasible. Similarly, the limitations of the steady state assumption for the length scales and uncertainties, which was necessary for reasons of available computation time, become apparent, in particular near the open boundary (station 1). Moreover, it has to be mentioned that the methodology proposed uses the covariance matrices in a time window in which the assump-tion of an invariant system is valid (i.e., February). For another period or month, this validity should be first controlled. This might require in practice a redefinition of a new Kalman gain for the new time window.

[41] An assessment of the overall forecast improvement

for a 3 h forecast and a 6 h forecast was also presented. The 3 h forecast and the 6 h forecast show higher hit rates than those of the model with no assimilation. In general, the hit rate of the forecast decreases in time as expected (i.e., the hit rate of 6 h forecast is lower than that of the 3 h forecast). The maximum hit rates are observed at the middle and lower layers because of the effect of vertical mixing at those layers. At the surface layer, the hit rates decrease rapidly in the forecasting process because of the dominant effect and variation of the wind field. On average, the direction hit rate of the model with no assimilation of 53% is increased to 69% for the 3 h forecast and 66% for the 6 h forecast. The magnitude and salinity hit rates of the model without assimilation of 65% and 64% respectively has increased to 84%, and 78% for the 3 h forecast and is almost maintained in the 6 h forecast (84%, and 75% respectively). The hit rates accuracy is operationally acceptable. However, the rapid decrease of the direction hit rates in the forecasting mode unlike the magnitude and salinity hit rates, indicates that the model capability for velocity components (i.e., north and east components) is not as accurate as the overall prediction capability for the salinity and current magnitude. [42] Finally, this paper shows that incorporating recent

measurement data through simple data assimilation techni-ques such as the steady state Kalman filter improves model outputs and forecasting abilities for operational purposes,

would like to thank Osaka Bay Regional Offshore Environmental Improve-ment Center for providing the data used in this research. Herman Gerritsen and Henk van den Boogaard of WL Delft Hydraulics are cordially thanked for their useful input and comments during the research and in the preparation of this manuscript.

References

Bertino, L., G. Evensen, and H. Wackernagel (2002), Combining geosta-tistics and Kalman filtering for data assimilation in an estuarine system, Inverse Probl., 18, 1 – 23.

Brasseur, P. (2006), Ocean data assimilation using sequential methods based on the Kalman filter, in Ocean Weather Forecasting: An Integrated View of Oceanography, edited by J. Verron, and E. Chassignet, pp. 271 – 316, Springer, Dordrecht, Netherlands.

Brasseur, P., et al. (2005), Data assimilation for marine monitoring and prediction: The MERCATOR operational assimilation systems and the MERSEA developments, Q. J. R. Meteorol. Soc., 131, 3561 – 3582. Cohn, S. E., and R. Todling (1996), Approximate data assimilation schemes

for stable and unstable dynamics, J. Meteorol. Soc. Jpn., 74, 63 – 75. Dee, D. P. (1991), Simplification of the Kalman filter for meteorological

data assimilation, Q. J. R. Meteorol. Soc., 117, 365 – 384.

El Serafy, G. Y., and A. E. Mynett (2004), Comparison of EKF and EnKF in Sobek River: Case study Maxau-IJssel, paper presented at 6th Inter-national Conference on Hydroinformatics, Int. Assoc. of Hydraul. Eng. and Res., Singapore.

El Serafy, G. Y., H. Gerritsen, A. E. Mynett, and M. Tanaka (2005), Im-provement of stratified flow forecasts in the Osaka Bay using the steady state Kalman filter, in Proceedings of the 31st IAHR Congress, edited by B.-H. Jun, et al., pp. 795 – 804, Int. Assoc. of Hydraul. Eng. and Res., Madrid.

El Serafy, G. Y., H. Gerritsen, A. E. Mynett, and M. Tanaka (2006), The use of Kalman filters in improving stratified flow forecasts for engineering purposes, in Proceedings of the 7th International Conference on Hydro-informatics, vol. 4, edited by P. Gourbesville, et al., pp. 2415 – 2422, Research, Chennai, India.

El Serafy, G. Y., H. Gerritsen, S. Hummel, A. H. Weerts, A. E. Mynett, and M. Tanaka (2007), Application of data assimilation in portable opera-tional forecasting systems—The DATools assimilation environment, Ocean Dyn., 57, 485 – 499.

Evensen, G. (1994a), Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics, J. Geophys. Res., 99(C5), 10,143 – 10,162.

Evensen, G. (1994b), Inverse methods and data assimilation in nonlinear ocean models, Physica D, 77, 108 – 129.

Evensen, G. (1997), Advanced data assimilation for strongly non-linear dynamics, Mon. Weather Rev., 125, 1342 – 1345.

Evensen, G. (2003), The ensemble Kalman filter: Theoretical formulation and practical implementation, Ocean Dyn., 53, 343 – 367.

Fukumori, I., and P. Malanotte-Rizzoli (1995), An approximate Kalman filter for ocean data assimilation: An example with an idealized Gulf Stream model, J. Geophys. Res., 100(C4), 6777 – 6793.

Jazwinsky, A. H. (1970), Stochastic Processes and Filtering Theory, Else-vier, New York.

Kalman, R. E. (1960), A new approach to linear filtering and prediction problems, J. Basic Eng., Ser. D, 82, 35 – 45.

Kalman, R. E., and R. S. Bucy (1961), New results in linear filtering and prediction theory, J. Basic Eng., Ser. D, 83, 95 – 108.

Kurapov, A. L., G. D. Egbert, R. N. Miller, and J. S. Allen (2002), Data assimilation in a baroclinic coastal ocean model: Ensemble statistics and comparison of methods, Mon. Weather Rev., 130, 1009 – 1025.

(19)

Kurapov, A. L., J. S. Allen, G. D. Egbert, R. N. Miller, P. M. Kosro, M. D. Levine, T. Boyd, and J. A. Barth (2005), Assimilation of moored velocity data in a model of coastal wind-driven circulation off Oregon: Multi-variate capabilities, J. Geophys. Res., 110, C10S08, doi:10.1029/ 2004JC002493.

Lesser, G. R., J. A. Roelvink, J. A. T. M. van Kester, and G. S. Stelling (2004), Development and validation of a three-dimensional morphologi-cal model, Coastal Eng., 51, 883 – 915.

Morf, M., G. S. Sidhu, and T. Kailath (1974), Some new algorithms for recursive estimation in constant, linear, discrete-time systems, IEEE Trans. Autom. Control, 19(4), 315 – 323.

Osaka Bay Regional Offshore Environmental Improvement Center, (2002), Report of preliminary environmental impact assessment on construction of ‘‘New Island’’ (in Japanese), Rep. 3, Osaka, Japan.

Pham, D. T., J. Verron, and M. C. Roubaud (1997), Singular evolutive Kalman filter with EOF initialization for data assimilation in oceanogra-phy, J. Mar. Syst., 16, 323 – 340.

Pham, D. T., J. Verron, and L. Gourdeau (1998), A singular evolutive Kalman filter for data assimilation in oceanography, C. R. Acad. Sci., Ser. II, 326, 255 – 260.

Tanaka, M., S. Inagaki, and T. Adachi (2006), A current and turbidity forecasting system for the environmental control of construction activ-ities in a stratified sea, in Proceedings of the 7th International Confer-ence on Hydroinformatics, vol. 3, edited by P. Gourbesville, et al., pp. 1577 – 1584, Research, Chennai, India.

Verlaan, M., and A. W. Heemink (1997), Tidal flow forecasting using reduced rank square root filters, Stochastic Hydrol. Hydraul., 11, 349 – 368.



G. Y. H. El Serafy and A. E. Mynett, Strategic Research and Development, WL Delft Hydraulics, P.O. Box 177, NL-2600 MHDelft, Netherlands. (ghada.elserafy@wldelft.nl; arthur.mynett@wldelft.nl)

Cytaty

Powiązane dokumenty

Nie te komputery, w których kolorowy obraz zmienia się od dotknięcia palcem ekranu i dzięki którym można dowiedzieć się w pół minuty na przykład, że jedyny w USA

Finally, adjustments to the planning tools at tactical and operational level are recommended, to plan as much as possible according to this most economical flow of materials..

Wojtowicza jest też opracowany indeks polsko-łacińsko-grecki i indeks łacińsko-polski nazw własnych autorów cyto­ wanych oraz indeks polsko-łacińsko-grecki i

First of all, in a short period of time it is difficult to see significant quality changes on the level of energy productivity and dependence on electricity prices due to no

Zasadniczą przeszkodę w powrocie do polityki wolnohandlow ej, obok trudności obiektyw nych (jak np. katastrofalny brak w ęgla i surowców, co zmuszało· do

Jeśli bowiem głębokie pokrewieństwo doświadczenia przedstawień dzieł sztuki klasycznej z doświadczeniem święta wiązało się ściśle z tym, że ich celem

For larger amount of members (i.e. 50 and above), the standard deviation of the mean of the RMSE of the EnPF is smaller than both the EnPF and the PF, which proves that the

In [Rommelse et al., 2007] the performance of the RM was investigated for different magnitudes of model errors, different line search algorithms, differ- ent minimization algorithms