Log in

Multivariate Prediction of Total Water Storage Changes Over West Africa from Multi-Satellite Data

  • Published:
Surveys in Geophysics Aims and scope Submit manuscript

Abstract

West African countries have been exposed to changes in rainfall patterns over the last decades, including a significant negative trend. This causes adverse effects on water resources of the region, for instance, reduced freshwater availability. Assessing and predicting large-scale total water storage (TWS) variations are necessary for West Africa, due to its environmental, social, and economical impacts. Hydrological models, however, may perform poorly over West Africa due to data scarcity. This study describes a new statistical, data-driven approach for predicting West African TWS changes from (past) gravity data obtained from the gravity recovery and climate experiment (GRACE), and (concurrent) rainfall data from the tropical rainfall measuring mission (TRMM) and sea surface temperature (SST) data over the Atlantic, Pacific, and Indian Oceans. The proposed method, therefore, capitalizes on the availability of remotely sensed observations for predicting monthly TWS, a quantity which is hard to observe in the field but important for measuring regional energy balance, as well as for agricultural, and water resource management. Major teleconnections within these data sets were identified using independent component analysis and linked via low-degree autoregressive models to build a predictive framework. After a learning phase of 72 months, our approach predicted TWS from rainfall and SST data alone that fitted to the observed GRACE-TWS better than that from a global hydrological model. Our results indicated a fit of 79 % and 67 % for the first-year prediction of the two dominant annual and inter-annual modes of TWS variations. This fit reduces to 62 % and 57 % for the second year of projection. The proposed approach, therefore, represents strong potential to predict the TWS over West Africa up to 2 years. It also has the potential to bridge the present GRACE data gaps of 1 month about each 162 days as well as a—hopefully—limited gap between GRACE and the GRACE follow-on mission over West Africa. The method presented could also be used to generate a near-real-time GRACE forecast over the regions that exhibit strong teleconnections.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. African monsoon multidisciplinary analysis.

References

  • Ahmed M, Sultan M, Wahr J, Yan E, Milewski A, Sauck W, Becker R, Welton B (2011) Integration of GRACE (gravity recovery and climate experiment) data with traditional data sets for a better understanding of the timedependent water partitioning in African watersheds. J Geol 41(1): doi:10.1130/G31812.1

  • Ali A, Lebel T (2009). The Sahelian standardized rainfall index revisited. Int J Climatol :1705–1714. doi: 10.1002/joc

  • Boone A, Decharme B, Guichard F, de Rosnay P, Balsamo G, Beljaars A, Chopin F, Orgeval T, Polcher J, Delire C, Ducharne A, Gascoin S, Grippa M, Jarlan L, Kergoat L, Mougin E, Gusev Y, Nasonova O, Harris P, Taylor C, Norgaard A, Sandholt I, Ottlé C, Poccard-Leclercq I, Saux-Picart S, Xue Y (2009) The AMMA land surface model intercomparison project (ALMIP). Bull Am Meteorol Soc 90(12):1865–1880. doi:10.1175/2009BAMS2786.1

    Article  Google Scholar 

  • Chen JL, Wilson CR, Tapley BD, Longuevergne L, Yang ZL, Scanlon BR (2010) Recent La Plata basin drought conditions observed by satellite gravimetry. J Geophys Res 115(D22):1–12. doi:10.1029/2010JD014689

    Google Scholar 

  • Crétaux J-F, Jelinski W, Calmant S, Kouraev A, Vuglinski V, Bergé Nguyen M, Gennero M-C, Nino F, Abarca Del Rio F, Cazenave A, Maisongrande P (2011) SOLS: a lake database to monitor in near real time water level and storage variations from remote sensing data. J Adv Space Res :1497–1507. doi: 10.1016/j.asr.2011.01.004

  • Diatta S, Fink AH (2014) Statistical relationship between remote climate indices and West African monsoon variability. Int J Climatol. doi:10.1002/joc.3912

  • Döll P, Kaspar F, Lehner B (2003) A global hydrological model for deriving water availability indicators: model tuning and validation. J Hydrol 270(1–2):105–134. doi:10.1016/S0022-1694(02)00283-4

    Article  Google Scholar 

  • Douville H, Conil S, Tyteca S, Voldoire A (2006) Soil moisture memory and West African monsoon predictability: artifact or reality? Clim Dyn 28(7–8):723–742. doi:10.1007/s00382-006-0207-8

    Google Scholar 

  • Efron B (1979) Bootstrap methods: another look at the jackknife. Ann Stat 7:1–26

    Article  Google Scholar 

  • Flechtner F (2007) GFZ Level-2 processing standards document for level-2 product release 0004. GRACE 327–743, Rev. 1.0

  • Fleming K, Awange JL (2013) Comparing the version 7 TRMM 3B43 monthly precipitation product with the TRMM 3B43 version 6/6A and BoM datasets for Australia. Aust Meteorol Oceanogr J 63(3):421–426

  • Forootan E, Didova O, Schumacher M, Kusche J, Elsaka B (2014) Comparisons of atmospheric mass variations derived from ECMWF reanalysis and operational fields, over 2003 to 2011. J Geod 88(5):503–514. doi:10.1007/s00190-014-0696-x

    Article  Google Scholar 

  • Forootan E, Didova O, Kusche J, Löcher A (2013) Comparisons of atmospheric data and reduction methods for the analysis of satellite gravimetry observations. J Geophys Res Solid Earth 118. doi:10.1002/jgrb.50160

  • Forootan E, Awange J, Kusche J, Heck B, Eicker A (2012) Independent patterns of water mass anomalies over Australia from satellite data and models. J Remote Sens Environ 124:427–443. doi:10.1016/j.rse.2012.05.023

    Article  Google Scholar 

  • Forootan E, Kusche J (2013) Separation of deterministic signals, using independent component analysis (ICA). Stud Geophys Geod 57:17–26. doi:10.1007/s11200-012-0718-1

    Article  Google Scholar 

  • Forootan E, Kusche J (2012) Separation of global time-variable gravity signals into maximally independent components. J Geod 86(7):477–497. doi:10.1007/s00190-011-0532-5

    Article  Google Scholar 

  • Giannini A, Saravanan R, Chang P (2003) Oceanic forcing of Sahel rainfall on interannual to interdecadal time scales. Science 302:1027–1030

    Article  Google Scholar 

  • Giannini A, Biasutti M, Held IM, Sobel AH (2008) A global perspective on African climate. Clim Change 90:359–383. doi:10.1007/s10584-008-9396-y

    Article  Google Scholar 

  • Grippa M, Kergoat L, Frappart F, Araud Q, Boone A, de Rosnay P, Lemoine J-M, Gascoin S, Balsamo G, Ottle C, Decharme B, Saux-Picart S, Ramillien G (2011) Land water storage variability over West Africa estimated by gravity recovery and climate experiment (GRACE) and land surface models. Water Resour Res 47:W05549. doi:10.1029/2009WR008856

    Article  Google Scholar 

  • Güntner A, Stuck J, Döll P, Schulze K, Merz B (2007) A global analysis of temporal and spatial variations in continental water storage. Water Resour Res 43(W05416). doi:10.1029/2006WR005247

  • Hansen JW, Mason SJ, Sun L, Tall A (2011) Review of seasonal climate forecasting for agriculture in sub-Saharan Africa. Expl Agric 47(2):205–240. doi: 10.1017/S0014479710000876

  • Heim RR (2002) A review of twentieth-century drought indices used in the United States. Bull Am Meteor Soc 83:1149–1165

    Google Scholar 

  • Houborg R, Rodell M, Li B, Reichle R, Zaitchik BF (2012) Drought indicators based on model-assimilated gravity recovery and climate experiment (GRACE) terrestrial water storage observations. Water Resour Res 48. doi:10.1029/2011WR011291

  • Huffman G, Bolvin D (2012) TRMM and other data precipitation data set documentation. Mesoscale atmospheric Processes Laboratory, NASA Goddard Space Flight Center and Science Systems and Applications Inc.

  • Ilin A, Valpola H, Oja E (2005) Semiblind source separation of climate data detects El Niño as the component with the highest interannual variability. In: Proceedings of the international joint conference on neural networks (IJCNN 2005), Montréal, Québec, Canada, pp 1722–1727

  • Kaplan A, Kushni Y, Cane MA, Blumenthal MB (1997) Reduced space optimal analysis for historical data sets: 136 years of Atlantic sea surface temperatures. J Geogr Res 102(C13):27835–27860

    Google Scholar 

  • Koster RD, Dirmeyer PA, Guo Z, Bonan G, Chan E, Cox P, Gordon CT, Kanae S, Kowalczyk E, Lawrence D, Liu P, Lu C-H, Malyshev S, McAvaney B, Mitchell K, Mocko D, Oki T, Oleson K, Pitman A, Sud YC, Taylor CM, Verseghy D, Vasic R, Xue Y, Yamada T (2004) Regions of strong coupling between soil moisture and precipitation. Science 305(5687):1138–1140. doi:10.1126/science.1100217

    Article  Google Scholar 

  • Kusche J (2007) Approximate decorrelation and non-isotropic smoothing of time-variable GRACE-type gravity field models. J Geod 81(11):733–749. doi:10.1007/s00190-007-0143-3

    Article  Google Scholar 

  • Kusche J, Schmidt R, Petrovic S, Rietbroek R (2009) Decorrelated GRACE time-variable gravity solutions by GFZ, and their validation using a hydrological model. J Geod 83:903–913. doi:10.1007/s00190-009-0308-3

    Article  Google Scholar 

  • Long D, Scanlon BR, Longuevergne L, Sun AY, Fernando DN, Save H (2013) GRACE satellite monitoring of large depletion in water storage in response to the 2011 drought in Texas. Geophys Res Lett 40(13):3395–3401. doi:10.1002/grl.50655

    Article  Google Scholar 

  • Ljung L (1987) System identification: theory for the user. Prentice Hall, Englewood Cliffs, N.J. 2005GL023316

  • Mayer-Gürr T, Eicker A, Kurtenbach E (2010) ITG-GRACE 2010 unconstrained monthly solutions. http://www.igg.uni-bonn.de/apmg/

  • Mohino E, Rodríguez-Fonseca B, Mechoso CR, Gervois S, Ruti P, Chauvin F (2011) Impacts of the tropical Pacific/Indian Oceans on the seasonal cycle of the West African monsoon. J Clim 24:3878–3891. doi:10.1175/2011JCLI3988.1

    Article  Google Scholar 

  • Nahmani S, Bock O, Bouin M-N, Santamaría-Gómez A, Boy J-P, Collilieux X, Métivier L, Panet I, Genthon P, de Linage C, Wöppelmann G (2012) Hydrological deformation induced by the West African Monsoon: comparison of GPS, GRACE and loading models. J Geophys Res 117:B05409. doi:10.1029/2011JB009102

    Google Scholar 

  • Nicholson SE (2000) The nature of rainfall variability over Africa on time scales of decades to millenia. Glob Planet Change 26:137–158

    Article  Google Scholar 

  • Nicholson SE et al (2003) Validation of TRMM and other rainfall estimates with a high-density gauge dataset for West Africa. Part I: validation of GPCC rainfall product and pre-TRMM satellite and blended products. J Appl Meteorol 42:1337–1354

    Article  Google Scholar 

  • Omondi P, Awange JL, Ogallo LA, Okoola RA, Forootan E (2012) Decadal rainfall variability modes in observed rainfall records over East Africa and their relations to historical sea surface temperature changes. J Hydrol 464—-465:140–156. doi:10.1016/j.jhydrol.2012.07.003

    Article  Google Scholar 

  • Preisendorfer R (1988) Principal component analysis in meteorology and oceanography. Elsevier, Amsterdam, p 426, ISBN:0444430148

  • Reynolds RW, Rayne NA, Smith TM, Stokes DC, Wang W (2002) An improved in situ and satellite SST analysis for climate. J Clim 15:1609–1625

    Article  Google Scholar 

  • Reager JT, Famiglietti JS (2013) Characteristic mega-basin water storage behavior using GRACE. Water Resour Res 49:3314–3329. doi:10.1002/wrcr.20264

    Article  Google Scholar 

  • Redelsperger J-L, Thorncroft ChD, Diedhiou A, Lebel T, Parker DJ, Polcher J (2006) African monsoon multidisciplinary analysis: an international research project and field campaign. Bull Am Meteor Soc 87:1739–1746. doi:10.1175/BAMS-87-12-1739

    Article  Google Scholar 

  • Rietbroek R, Brunnabend SE, Dahle C, Kusche J, Flechtner F, Schröter J, Timmermann R (2009) Changes in total ocean mass derived from GRACE, GPS, and ocean modeling with weekly resolution. J Geophys Res 114(C11004):2009J. doi:10.1029/C005449

    Google Scholar 

  • Rietbroek R, Fritsche M, Dahle C, Brunnabend S-E, Behnisch M, Kusche J, Flechtner F, Schröter J, Dietrich R (2014) Can GPS-derived surface loading bridge a GRACE mission gap? Surv Geophys. doi:10.1007/s10712-013-9276-5

  • Rodell M, Houser PR, Jambor U, Gottschalck J, Mitchell K, Meng K, Arsenault C-J, Cosgrove B, Radakovich J, Bosilovich M, Entin JK, Walker JP, Lohmann D, Toll D (2004) The global land data assimilation system. Bull Am Meteorol Soc 85(3):381–394

    Article  Google Scholar 

  • Rodríguez-Fonseca B, Janicot S, Mohino E, Losada T, Bader J, Caminade C, Chauvin F, Fontaine B, García-Serrano J, Gervois S, Joly M, Polo I, Ruti P, Roucou P, Voldoire A (2011) Interannual and decadal SST-forced responses of the West African monsoon. Atmos Sci Lett 12(1):67–74. doi:10.1002/asl.308

    Article  Google Scholar 

  • Saji NH, Goswami BN, Vinayachandran PN, Yamagata T (1999) A dipole mode in the tropical Indian Ocean. Nature 401:360–363. doi:10.1038/43854

    Google Scholar 

  • Scanlon BR, Faunt CC, Longuevergne L, Reedy RC, Alley WM, McGuire VL, McMahon PB (2012) Groundwater depletion and sustainability of irrigation in the US High Plains and Central Valley. Proc Natl Acad Sci USA 109(24):9320–9325. doi:10.1073/pnas.1200311109

    Article  Google Scholar 

  • Schmidt R, Flechtner F, Meyer U, Neumayer K-H, Dahle Ch, König R, Kusche J (2008) Hydrological signals observed by the GRACE satellites. Surv Geophys 29(4–5):319–334

    Article  Google Scholar 

  • Schmidt R, Petrovic S, Güntner A, Barthelmes F, Wünsch J, Kusche J (2008) Periodic components of water storage changes from GRACE and global hydrology models. J Geophys Res 113:B08419

    Google Scholar 

  • Schumacher M, Eicker A, Kusche J, Schmied HM, Döll P (2014) Covariance analysis and sensitivity studies for GRACE assimilation into WGHM. IAG Scientific Assembly Proceedings 2014 (in press)

  • Schuol J, Abbaspour KC, Yang H, Srinivasan R, Zehnder AJB (2008) Modeling blue and green water availability in Africa. Water Resour Res 44. doi:10.1029/2007WR006609

  • Schuol J, Abbaspour KC (2006) Calibration and uncertainty issues of a hydrological model (SWAT) applied to West Africa. Adv Geosci 9:137–143

    Article  Google Scholar 

  • Speth P, Christoph M, Diekkrüger B (2011) Impacts of global change on the hydrological cycle in West and Northwest Africa. Speth, Peter; Christoph, Michael; Diekkrüger, Bernd (Eds.). Springer, Berlin Heidelberg, p 675. ISBN:3642129560

  • Tapley B, Bettadpur S, Watkins M, Reigber C (2004) The gravity recovery and climate experiment: mission overview and early results. Geophys Res Lett 31. 10.1029/2004GL019920

  • van Dijk AIJM, Peña-Arancibia JL, Sheffield J, Beck HE (2013) Global analysis of seasonal streamflow predictability using an ensemble prediction system and observations from 6192 small catchments worldwide. Water Resour Res 49(5):2729–2746. doi:10.1002/wrcr.20251

    Article  Google Scholar 

  • von Storch H, Navarra A (1999) Analysis of climate variability. Springer, p 342, ISBN 978-3-540-66315-7

  • Wahr J, Molenaar M, Bryan F (1998) Time variability of the earth’s gravity field: hydrological and oceanic effects and their possible detection using GRACE. J Geophys Res 103(B12):30205–30229. doi:10.1029/98JB02844

    Article  Google Scholar 

  • Wang K, Dickinson RE (2012) A review of global terrestrial evapotranspiration: observation, modeling, climatology, and climatic variability. Rev Geophys 50(2). doi:10.1029/2011RG000373

  • Werth S, Güntner A, Schmidt R, Kusche J (2009) Evaluation of GRACE filter tools from a hydrological perspective. Geophys J Int 179:1499–1515. doi:10.1111/j.1365-246X.2009.04355.x

    Article  Google Scholar 

  • Werth S, Güntner A, Petrovic S, Schmidt R (2009) Integration of GRACE mass variations into a global hydrological model. Earth Planet Sci Lett 277(1):166–173

    Article  Google Scholar 

  • Westra S, Brown C, Lall U, Sharma A (2007) Modeling multivariable hydrological series: principal component analysis or independent component analysis? Water Resour Res 43(6):W06429. doi:10.1029/2006WR005617

    Article  Google Scholar 

  • Westra S, Sharma A, Brown C, Lall U (2008) Multivariate streamflow forecasting using independent component analysis. Water Resour Res 44(2):W02437. doi:10.1029/2007WR006104

    Article  Google Scholar 

  • **e H, Longuevergne L, Ringler C, Scanlon BR (2012) Calibration and evaluation of a semi-distributed watershed model of Sub-Saharan Africa using GRACE data. Hydrol Earth Syst Sci 16(9):3083–3099. doi:10.5194/hess-16-3083-2012

    Article  Google Scholar 

  • Zaitchik BF, Rodell M, Reichle RH (2008) Assimilation of GRACE terrestrial water storage data into a land surface model: results for the Mississippi River Basin. J Hydrometeor 9(3):535–548. doi:10.1175/2007JHM951.1

    Article  Google Scholar 

Download references

Acknowledgments

The authors would like to thank M. J. Rycroft (Editor in Chief) and anonymous reviewers for their useful comments, which considerably improved this paper. We also thank S. Nahmani (LAboratoire de Recherche en Géodésie, France) for his detailed comments on the earlier version of this study. We are grateful for the GRACE, WGHM, TRMM, and SST data, as well as climate indices used in this study. E. Forootan and J. Kusche are grateful for the supports by the German Research Foundation (DFG), under the project DFG BAYES-G. The Ohio State University component of the research is supported by the NASA’s Advanced Concepts in Space Geodesy Program (Grant No. NNX12AK28G) and by the Chinese Academy of Sciences/SAFEA International Partnership Program for Creative Research Teams (Grant No. KZZD-EW-TZ-05). The authors are grateful for the data used in this study. This is a TIGeR Publication no. 510.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ehsan Forootan.

Appendices

Appendix 1: Computational Details of Total Water Storage Fields

In order to prepare the data sets for analysis, the following processing steps were applied.

  • The GRACE Level-2 data that are used here are derived in terms of fully normalized spherical harmonic (SH) coefficients of the geopotential fields (Flechtner 2007). Firstly, the fields were augmented by the degree-1 term from Rietbroek et al. (2009) in order to include the variation of the Earth’s center of mass with respect to a crust-fixed reference system.

  • GRACE SHs at higher degrees are affected by correlated noise and are, therefore, smoothed by applying the DDK2 decorrelation filter (Kusche et al. 2009). Werth et al. (2009a) found that the DDK2-filtered GRACE solutions are generally in good agreement with the output of global hydrological models. However, GRACE solutions are also contaminated by errors due to incomplete reduction of short-term mass variations by de-aliasing models (Forootan et al. 2013, 2014). We found that the impact of atmospheric de-aliasing errors on the GRACE-derived TWS over West Africa is negligible (see atmospheric errors over the Niger Basin in Forootan et al. 2014).

  • GRACE DDK2 filtered solutions up to degree and order 120 were then used to generate the global TWS values according to the approach of Wahr et al. (1998).

  • Similar to the GRACE products above, the DDK2 filter was applied to the gridded WGHM-TWS data set in order to preserve exactly the same spectral content as with the filtered GRACE products.

  • After filtering, all data sets were converted to 0.5° × 0.5° grids similar to the WGHM-TWS outputs.

  • From each data set, a rectangular region that includes West Africa (latitude between 0° and 25°N and longitude between −20° and 10°E) was selected.

Lake Volta (see Fig. 7) is one of the largest man-made reservoirs in the world, created by the Akosombo Dam, which holds back the water for generating hydroelectric power (for details see Speth et al. 2011). Satellite altimetry observations indicate a sharp increase of water level since mid 2007, where much of the excess water resulted from heavy rainfall within the catchment (Crétaux et al. 2011). This introduces an artificial TWS anomaly located over the lake, which is removed to avoid its misinterpretation as a part of subsurface TWS changes. The equivalent water height (EWH) change of Volta was computed by assuming a grid mask representing a unit change in EWH of 1 mm over the entire lake surface and zero elsewhere. The grid mask has been converted into a set of spherical harmonic coefficients up to degree 120 and subsequently filtered using the same DDK2 filter used for filtering the original GRACE-TWS data. Then, each field was scaled using the lake height time series (in mm) derived from the results of Crétaux et al. (2011). The averaged storage changes derived from GRACE-TWS (from GFZ) and altimetry are shown in Fig. 7. Both altimetry and GRACE GFZ-TWS indicate an increase of water storage within the lake. The amplitude of the signal derived from GRACE GFZ-TWS is larger than that of the altimetry likely since GRACE-TWS also reflects the groundwater signal of the surrounding area of the lake. For the lake area, we estimate a TWS increase of 2.95 ± 1.32 km3 year−1, during 2003 to 2010. The time series of Lake Volta water storage changes were then removed from GRACE-TWS fields (including both GFZ and ITG2010).

Fig. 7
figure 7

Overview of water storage changes of Volta Lake. The graph on top shows the location of the lake, while that of the bottom-left compares the averaged contribution of Volta Lake level changes (derived from altimetry in black) with the averaged TWS variations derived from GRACE (GFZ-TWS products, in red). The bottom-right graph shows the GRACE GFZ-TWS signal after removing the water storage signal of Volta Lake

In order to compare the signal strength over the region, the signal root mean square (RMS) value and the linear trend of the three mentioned TWS data sets (GFZ, ITG2010, and WGHM) are computed for the period January 2003–August 2009, in which the three data sets were available (see Fig. 8). From the RMS, one concludes that all the three data sets show a strong variability over the tropical and the Gulf of Guinea coastal regions. The computed linear trends, however, are not consistent. Particularly, GRACE-derived TWS changes show a mass gain over Volta Lake, which we remove from the GRACE-TWS fields before performing decomposition. Removing such artificial anomaly is necessary, since otherwise the amplitude of TWS forecast over the lake will be overestimated.

Fig. 8
figure 8

Comparing the signal variability (RMS) and linear trends of TWS data used in this study after smoothing using the Kusche et al. (2009)’s DDK2 filter. a) TWS data of GRACE GFZ RL04, b TWS data of GRACE ITG2010, and c TWS of WGHM

Appendix 2: Details of ICA and ARX Methods

This appendix provides details of computations regarding to the methodology described in Sect. 3.

1.1 The ICA Computations

ICA decomposition is performed here by applying a 2-step algorithm (Forootan and Kusche 2012) on the available data sets, where step 1 consists of data decorrelation using principal component analysis (PCA). In step 2, the \(j\)-dominant components of PCA are rotated to be as independent from each other as possible. Storing the available data in a \(n \times p\) data matrix \(\mathbf{X}\), after removing their temporal mean, where \(n\) is the number of months and \(p\) is the number of grid points, ICA decomposes \(\mathbf{X}\) as

$$\begin{aligned} \mathbf{X}\simeq \mathbf{X}_j{= {\bar{\mathbf{P}}_j \mathbf{R}_j \mathbf{\Lambda }_j \mathbf{R}_j^T\bar{\mathbf{E}}_j^T}}. \end{aligned}$$
(4)

In Eq. (4), \(\bar{\mathbf{P}}_j \mathbf{\Lambda }_j \bar{\mathbf{E}}_j^T\) is derived from the PCA decomposition of \(\mathbf{X}\) in step 1. Therefore, \(\mathbf{\Lambda }_j\) is an \(j \times j\) diagonal matrix that stores the singular values arranged with respect to the magnitude, \(\bar{\mathbf{E}}_j\) (\(j \times p\)) contains the corresponding unit-length spatial eigenvectors, \(\bar{\mathbf{P}}_j\) (\(n \times j\)) contains the associated normalized temporal components, and \(j<n\) is the number of retained dominant modes (Preisendorfer 1988). The orthogonal rotation matrix \(\hat{\mathbf{R}}_j\) (\(j \times j\)) is defined in step 2, so that it rotates PCs and make them as statistically independent as possible. The method equals to temporal ICA (Forootan and Kusche 2012), which is simply called ICA in the paper. Considering Eqs. (1) and (2), \(\mathbf{Y}\) and \(\mathbf{U}\) are equivalent to \({\bar{\mathbf{P}}}\mathbf{R}\), while \(\mathbf{A}\) and \(\mathbf{B}\) are equivalent to \(\mathbf{\Lambda } {\bar{\mathbf{E}}} \mathbf{R}\). An optimum \(\mathbf{R}\) was found by digonalization of the fourth-order cross cumulants of the dominant temporal components \(\bar{\mathbf{P}}\) (see details in Forootan and Kusche 2012).

For properly selecting the subspace dimension \(j\) or \(j'\), we used a Monte Carlo approach which simulates data from a random distribution \({\mathbf{N}(\mathbf 0,\mathbf \Sigma )}\), with \(\mathbf{\Sigma }\) containing the column variance of \(\mathbf{X}\). The null hypothesis is that \(\mathbf{X}\) is drawn from such a distribution (see also Preisendorfer 1988, pp. 199–205). To apply the rule, 100 time series realizations of \({\mathbf{N}(\mathbf 0,\mathbf \Sigma )}\) are generated, their eigenvalues computed and placed in decreasing order. The 95th and 5th percentile of the cumulative distribution are then plotted (red lines in Fig. 9). Eigenvalues from the actual data sets that are above the derived confidence boundaries are unlikely to result from a data set consisting of only noise. To estimate the uncertainties of the eigenvalues, we randomly selected a subsample of \(\mathbf{X}\) and applied PCA, then selected another subsample and repeated this operation 200 times. This approach follows the ‘bootstrap**’ method as presented, e.g., in Efron (1979) and yields uncertainty estimates (see error bars in Fig. 9). The repeat number of 200 is chosen experimentally to be sure that the distribution of the estimated eigenvalues is independent from the selections of the subsamples.

To illustrate what we describe above, Fig. 9 shows the eigenvalue spectrum of the centered time series of GRACE GFZ-derived TWS, SST and rainfall computed using PCA. The significance levels are shown by red lines and the error bars show the uncertainties of eigenvalues. The eigenvalues above the red lines are statistically significant. The significant eigenvalues along with their orthogonal components are rotated toward independence using Eq. (4) and interpreted in Sect. 4.

Fig. 9
figure 9

Eigenvalue results derived from implementing the PCA method on a time series of GRACE GFZ-TWS maps, b maps of SST over the Atlantic, c maps of SST over the Pacific, d maps of SST over the Indian Ocean basins, and e West African rainfall maps from TRMM. Uncertainties are shown by error bars around each eigenvalues. The red lines correspond to the significant test, described in ‘Appendix 2.’ The variance fractions of the dominant eigenvalues are represented in each graph

Based on the uncertainties of the PCA results (Fig. 9), in order to estimate the uncertainty of the ICs [Eq. (4)], we generated 100 realizations of \(\mathbf{X}\), reconstructed by \(\bar{\mathbf{P}}_j, \mathbf{\Lambda }_j\), and \({\bar{\mathbf{E}}_j}\) along with 100 realizations of their errors. Then, applying Eq. (4) to the realizations allows the estimation of uncertainties (see, e.g., error bars in Figs. 1, 2, 3).

The projection of the data \(\mathbf{X}\) onto the i’th spatial pattern of the ICA \(\hat{\mathbf{p}}_i=\mathbf{X}\hat{\mathbf{e}}_i\), provides its corresponding temporal evolution

$$\begin{aligned} \hat{\mathbf{p}}_i(t) = \sum _{s=1}^{p} {x}(t,s)\;\hat{\mathbf{e}}_i(s), \end{aligned}$$
(5)

where \(t\) is time (\(1,\ldots ,n\)) and \(s\) is the number of grid points (\(1,\ldots ,p\)).

1.2 The ARX Computations

Considering Eq. (3) as the ARX model, the ARX forecast requires two steps: (i) The coefficients \((a_1,\ldots ,a_{n_a})\) and \((b_{q,1},b_{q,2},\ldots ,{b_{q,{n_b}}}), q=1,\ldots , m\) are estimated, e.g., using a least squares approach. This step is usually referred to as ‘simulation’ or ‘training step’ in the literature (see, e.g., Ljung 1987). Step (i) is performed under the assumption that the output and inputs up to the time \(t=t_n-1\) are known. Furthermore, the outputs and exogenous values on the right-hand side of Eq. (3) are not stochastic. To avoid negative indices, one might consider the observations \(\mathbf{y}(t) = [{y}(t),{y}(t-1),\ldots ,{y}(c)]^T\), where \(c= {\rm max}(n_a,n_b)+ {\rm max}(k_q)+1\). Eq. (3) is expanded as

$$\begin{aligned} \mathbf{y} =\left[ {\begin{array}{cccccc} -y(t-1)&\cdots&-y(t-n_a)&{u}_q(t-k_q)&\cdots&{u}_q(t-k_q-n_b+1)\\ -y(t-2)&\cdots&-y(t-n_a-1)&{u}_q(t-k_q-1)&\cdots&{u}_q(t-k_q-n_b)\\ \vdots&\vdots&\vdots&\vdots&\vdots&\vdots \\ -y(c-1)&\cdots&-y(c-n_a)&{u}_q(c-k_q)&\cdots&{u}_q(c-k_q-n_b+1)\\ \end{array}}\right]\left[ \begin{array}{c} a_1 \\ \vdots \\ a_{n_a}\\ b_{q,1}\\ \vdots \\ b_{q,n_b} \end{array}\right]+\varvec{\** }(t), \end{aligned}$$
(6)

where \(q=1,\ldots ,m\) and \(\varvec{\** }(t) = [{\xi }(t), {\xi }(t-1),\ldots , {\xi }(c)]^T\). Eq. (6) can be rewritten compactly as

$$\begin{aligned} {\mathbf{y}(t)} = \mathbf{\Phi }(t) \varvec{\Theta }+\varvec{\** }(t). \end{aligned}$$
(7)

The least squares estimation of the unknown coefficients is derived from

$$\begin{aligned} \hat{\varvec{\Theta }} = {\left( {\mathbf{\Phi }(t)}^T{\mathbf{\Phi }(t)}\right) }^{-1}{\mathbf{\Phi }(t)}^T\; {\mathbf{y}(t)}. \end{aligned}$$
(8)

The quality of the fit (\(\eta\)) can be assessed by computing the signal-to-noise ratio as

$$\begin{aligned} \eta = 1-\frac{{\mathbf{y}(t)}^T\mathbf{\Phi }(t)\hat{\varvec{\Theta }}}{{\mathbf{y}(t)}^T\mathbf{y}(t)}. \end{aligned}$$
(9)

The residual of the ARX model (\(\hat{\varvec{\** }}(t) = [\hat{\xi }(t),\hat{\xi }(t-1),\ldots ,\hat{\xi }(c)]^T\)) can be estimated as

$$\begin{aligned} \hat{\varvec{\** }}(t) = \mathbf{y}(t)-{\mathbf{\Phi }(t)}\;\hat{\varvec{\Theta }}. \end{aligned}$$
(10)

In step (ii), based on \(\hat{\varvec{\Theta }} = \left[ \hat{a}_1 \; \ldots \; \hat{a}_{n_a} \; \hat{b}_{q,1} \; \hat{b}_{q,2} \; \ldots \; \hat{b}_{q,{n_b}} \right] ^T\), when the inputs \({u}_q(t)\) are known, one can forecast the output \(\hat{y}(t_n)\) at time \(t_n\) using

$$\begin{aligned} \hat{y}(t_n)=-\sum _{i=1}^{n_a}a_{i}y(t_n-i)+\sum _{q=1}^{m} \sum _{l=1}^{n_b}{b_{q,l}{u}_q(t_n-k_q-(l-1))}. \end{aligned}$$
(11)

To estimate the uncertainty of the ARX simulation, using Monte Carlo sampling, we numerically generate several realizations of the ICs (described before). By inserting them into Eq. (3) and fitting ARX models, we are able to perform an error assessment of the fitted model up to the time \(t_n\). For error estimation of the forecast (the ARX value at the time \(t_n+1\) and later), however, one should compute an accumulated error, since there is no observed value for the output \(y\) at time \(t_n+1\) and later.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Forootan, E., Kusche, J., Loth, I. et al. Multivariate Prediction of Total Water Storage Changes Over West Africa from Multi-Satellite Data. Surv Geophys 35, 913–940 (2014). https://doi.org/10.1007/s10712-014-9292-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10712-014-9292-0

Keywords

Navigation