Estimating global radiation at ground level from satellite images Knut-Frode Dagestad DOCTOR SCIENTIARUM THESIS IN METEOROLOGY AT UNIVERSITY OF BERGEN MAY 2005 Preface This synthesis and collection of papers constitute a thesis presented in partial fulfillment of the requirement for the degree of Doctor Scientiarum in meteorology at the Geophysical Institute, University of Bergen, Norway. This thesis discusses the Heliosat algorithm which estimates solar radiation at ground level from satellite images. The performance of various versions of the algorithm has been analysed, and modifications are suggested. Acknowledgments First of all I want to thank my two supervisors, Arvid Skartveit and Jan Asle Olseth for excellent support and invaluable feedback. This work is a part of the project "Heliosat-3", which has been funded by the European Commision. It has been a pleasure to work in the Heliosat-3 team, and the mood has always been humoristic and positive, despite that we were missing our satellite for a long time. A special thanks goes to the German colleagues Richard Müller, Rolf Kuhlemann and Hermann Mannstein who invited me to stay in their private homes during three of my stays in Germany. I also want to thank all the staff and students at the Geophysical Institute for a very good working environment. In particular I want to mention Børge and Gard with whom I have shared my refrigerator and many enjoyable moments. I also want to mention Yngvar Gjessing for sharing some of his wisdom in many fields. Many thanks to Christiane and Frode for proofreading the thesis. I am also very grateful to the Nansen Center for financially supporting this work the last nine months, and also for the very interesting new job which I am looking forward to do wholeheartedly from now on. Finally, thanks to my family and friends for helping me having a somewhat normal social life (at least until recent months). Bergen, May 2005 Knut-Frode Dagestad Table of Contents 1 Introduction...................................................................................................................................... .1 2 History of the Heliosat algorithm..................................................................................................... .2 2.1 The original version.................................................................................................................. .2 2.2 Heliosat-1...................................................................................................................................3 2.3 Heliosat-2...................................................................................................................................4 2.4 Heliosat-3 (objectives of this thesis)..........................................................................................5 3 The concepts of the Heliosat algorithm.............................................................................................6 3.1 An empirical approach...............................................................................................................6 3.2 A physical approach...................................................................................................................6 3.3 A sensitivity analysis................................................................................................................ .8 4 Summary of results..........................................................................................................................12 4.1 Paper I...................................................................................................................................... 12 4.2 Paper II.....................................................................................................................................13 4.3 Paper III................................................................................................................................... 14 4.4 Paper IV................................................................................................................................... 14 4.5 Paper V.................................................................................................................................... 15 5 Concluding remarks.........................................................................................................................15 6 Bibliography.................................................................................................................................... 16 7 Part II - The papers I-V................................................................................................................... 17 List of Papers Paper I Dagestad, K-F. (2004). Mean bias deviation of the Heliosat algorithm for varying cloud properties and sunground-satellite geometry. Theoretical and Applied Climatology, 79, 215–224. Paper II Müller, R. W., Dagestad, K-F., Ineichen, P., Schroedter, M., Cros, S., Dumortier, D., Kuhlemann, R., Olseth, J. A., Piernavieja, C., Reise, C., Wald, L. and Heinemann, D. (2004). Rethinking satellite based solar irradiance modelling - The SOLIS clear sky module. Remote Sensing of the Environment, 91, 160-174. Paper III Dagestad, K-F. and Olseth, J.A. (2005). An alternative algorithm for calculating the cloud index. Manuscript. Paper IV Dagestad, K-F. and Wald, L. (2005). Assessment of the database HelioClim-2 of hourly irradiance derived from satellite observations. Draft. Paper V Dagestad, K-F. (2005). Simulations of bidirectional reflectance of clouds with a 3D radiative transfer model. Manuscript. 1 Introduction Every second the sun radiates more energy than people have used since the beginning of time. The amount of solar energy reaching earth is 1.76 x 1017 joules per second; more than 10000 times the global energy consumption today. Still this enormous source of energy accounts for only 0.1% of the total consumption, whereas 77% comes from fossil fuels (Worldwatch Institute 2003). The reasons for the low exploitation rate of solar energy are that it is difficult to collect since it is spread over the whole earth, and difficult to predict since it is fluctuating in time and space. To increase the efficiency of solar thermal power plants a detailed knowledge of the spatial and temporal variation of solar irradiance is needed. Such a climatology can be made by interpolating between measurement stations, as has been done in e.g. the European Solar Radiation Atlas, ESRA (Scharmer 1994). However, low spatial and temporal resolution of such data has led to nonoptimal site selection and incorrect system sizing, and thus unnecessary use of conventional energy sources. Over the last two decades satellite-based retrieval of solar radiation at ground level has proven to be highly valuable for the solar energy community. With satellite pixel sizes of typically 2.5 kilometres, the spatial resolution of the estimates is much better than the data interpolated from ground measurements. It has also been found that for hourly values of global irradiance satellite retrievals are more accurate than interpolating ground measurements from stations which are more than 30 kilometres apart (Zelenka et al. 1999). However, the global markets for renewable energy sources such as solar and wind power are expected to see a dramatic expansion in the near future (Worldwatch Institute 2003), and there is a demand for solar radiation data of even higher quality. This can be made possible with more sophisticated satellite technologies and by improvement of the algorithms for conversion of satellite data into solar radiation data. Heliosat is an algorithm which has been developed to estimate global horizontal irradiance at ground level from images taken in the visible range by the European meterological satellite series Meteosat. Starting with the launch of Meteosat-8 in August 2002 these satellites have now increased capabilities; the size of a pixel is now 1 kilometre, compared to 2.5 kilometres earlier, and the temporal resolution is increased from 2 to 4 images per hour. In addition, the number of spectral channels is increased from 3 to 12, making it possible to get a more accurate description of the atmospheric state. The work of this thesis deals with the improvement of the Heliosat algorithm, partly by taking advantage of the enhanced capabilities of the new generation of the Meteosat satellites. The thesis is composed of two parts: Part I provides an overview of the thesis. This introduction is the first chapter of this part. Chapter 2 gives an overview of the history of the Heliosat algorithm, and describes the various versions of the algorithm which are referred to later in the thesis. Section 2.4 discusses the prospects of Heliosat-3, the latest version which has been developed within the EU-project of the same name, based partly on the work of this thesis. Therefore this section also gives the objectives of the thesis. Chapter 3 discusses the concepts behind the Heliosat algorithm, and also gives an analysis of how sensitive the output of the algorithm is to some of its components. A summary of the five papers of the thesis is given in chapter 4 and concluding remarks are given in chapter 5. Part II encompasses the five papers which constitute the main scientific work of the thesis. 1 2 History of the Heliosat algorithm The first experimental weather satellite, TIROS-I, was launched by the USA in 1960. After several years of experiments the satellites were gradually improved and better adapted to their specific uses. 17 years later, in 1977, the European Space Agency launched Meteosat-1, the first satellite of the European meteorological satellite system. The main purpose of the Meteosat-satellites was to improve weather forecasts by giving the meteorologists a visual overview of the cloud cover on a global scale. In addition, several other applications of the satellite images quickly emerged; among them were methods to estimate the solar irradiance at ground level. However, the satellite data were very simple; each pixel of the images consisted of a digital count between 0 and 255, and these pixel counts could not even be reliably calibrated into radiances. Despite the input being simple, the output of these algorithms was surprisingly accurate when compared to ground measurements. The Heliosat algorithm, originally proposed by Cano et al. (1986), was one of the most popular algorithms because it was accurate and easy to implement (Grüter et al. 1986). Heliosat was widely used in operational schemes around the world (Wald et al. 1992), and over the years it was modified several times. The naming of the different versions can be confusing, as a lot of versions exist with only minor differences from others. The following sections describe the versions which are referred to later in the papers of this thesis, and at the same time they give an overview of the major steps of the evolution of the Heliosat algorithm. 2.1 The original version The first journal paper about the Heliosat algorithm was published in 1986 (Cano et al. 1986). This original version used uncalibrated counts of the Meteosat High Resolution Visible (HRV) sensor to calculate a reflectivity of the pixels: = C G clear (1) Here C is the digital counts of a pixel, an 8-bit number between 0 and 255, and Gclear is the clear sky global irradiance at ground level from an empirical model. The clear sky model in the first Heliosat scheme was very simple as it used only the solar elevation as input, and no information about the atmospheric turbidity at the given site. As a second step a cloud index n was calculated from a time series of reflectivities: n= −clear cloud −clear (2) Here ρclear and ρcloud are the reflectivities corresponding to clear and overcast conditions, respectively. ρclear was decided by a histogram technique so that it is the "most frequent low reflectivity" of a given pixel for a given month. Similarly ρcloud was chosen as the "most frequent high reflectivity". A simple linear relation was then assumed between the cloud index and the clearness index kc: k c≡ G =anb G TOA (3) The clearness index is the ratio of the actual global irradiance, G, to the irradiance at the top of the atmosphere, GTOA. The parameters a and b were tuned to ground measurements at a number of sites to minimise the deviation. Different values of the parameters were found for different sites and 2 times of day, reflecting diurnal and spatial variation of atmospheric turbidity. Although this scheme was very simple in its principle, implementation was relatively complex: the parameters a and b varied between the sites, and interpolation and extrapolation was done to provide global application. In some operational versions different values of these parameters were also used for morning, noon and afternoon. Consequently the accuracy was good at sites and times for which the constants a and b were tuned, and less good at sites where they were interpolated. 2.2 Heliosat-1 "Heliosat-1" refers in this thesis to the modified version of the algorithm which was developed within the EU-project "Satel-Light" (Fontoynont et al., 1998). This was the first operational large scale implementation of the algorithm, and global irradiance and derived products for the period 1996-2000 are disseminated for most of Europe on the web server www.satel-light.com. The main differences from the original version are described in Beyer et al. (1996), Fontoynont et al. (1998) and Hammer (2000): The reflectivity (ρ) is now calculated with: = C −C atm G TOA (4) The radiation scattered back to the satellite from atmospheric molecules, Catm, is subtracted from the satellite measurements so that ρ is a reflectivity of the ground and clouds only. An empirical expression for Catm was developed by Beyer et al. (1996) and later modified by Hammer (2000). The expression was tuned to digital counts of cloud free pixels over sea, assuming that the reflected radiance from the sea surface is negligible compared to the scattered radiance from air molecules. Instead of normalising with a modelled clear sky irradiance, the irradiance at the top of the atmosphere, GTOA, is now used in the normalisation. The clearness index kc (Equation 3) is replaced by the clear sky index k, the actual global irradiance, G, divided by the output of a clear sky model, Gclear: k≡ G G clear (5) In contrast to the relation between the cloud index and the clearness index (Equation 3) which had to be tuned to ground data at all sites, the relation between the cloud index and the clear sky index is now the same for any site. This new empirical relation is given by: 1.2 1−n k= 2 2.0667−3.6667 n1.6667 n 0.05 for n −0.2 for n ∈[−0.2, 0.8] for n ∈[ 0.8,1.1] for n1.1 (6) The cloud index is still calculated with Equation 2. In the original version the spatial and temporal variation of atmospheric turbidity was accounted for indirectly by tuning the parameters of the relation between the cloud index and the clearness index (Equation 3). In the Heliosat-1 version the atmospheric turbidity is a directly input parameter to the clear sky model. This is more convenient because such turbidity parameters are already available 3 for other purposes, and they are also easier to interpret physically, in contrast to the parameters a and b of Equation 3. The clear sky model used in the Satel-Light project consists of a model for the direct irradiance from Page (1996) and a model for the diffuse irradiance from Dumortier (1995). As input they used monthly values of Linke turbidities from a database developed by Dumortier (1998), height above sea level and solar elevation. The Linke turbidity coefficients are commonly used in meteorology, and account for the combined attenuation of broadband solar irradiance by aerosols and water vapour. For the operational scheme an average of 3 pixels in the north-south direction and 5 pixels in the east-west direction was used, corresponding to a roughly square area in Europe where pixels are longer in the north-south direction due to the oblique viewing angle. The averaging led to better results, probably because the scale of this larger pixel-cluster better correlates the interval between subsequent images (30 minutes) with typical movement of clouds which are obstructing the path between the ground and the satellite sensor. Besides, this averaging is smoothing out the variability related to non-lambertian reflectivity, an issue which is discussed in Paper V of this thesis. In Heliosat-1 the ground albedo was calculated separately for each slot (images acquired at the same time (UTC) of day belong to the same slot) for each month. This gave a better correlation with ground measurements, probably because the sun-ground-satellite configuration was kept fairly constant, thus minimising the effect of non-lambertian reflectivity. 2.3 Heliosat-2 Heliosat-2 uses the same principles as Heliosat-1, but a major difference is that calibrated radiances instead of raw digital counts are used as input (Rigollier et al. 2004). The HRV sensor of Meteosat is not calibrated routinely by Eumetsat, but a method for operational calibration was developed and used at Ecole de Mines in France (Lefevre et al. 2000, Rigollier et al. 2002). Heliosat-2 was also developed at Ecole des Mines, mainly during the EU-project SoDa from 2000-2001 (www.sodais.com). With calibrated radiances as input, Heliosat-2 uses the opportunity to replace some of the empirical parameters in the scheme with known physical/empirical models from external sources: The correction for the backscattered radiation from the atmosphere (Equation 4) is based on the ESRA clear sky model (Rigollier et al. 2000, Geiger et al. 2002). Here the clear sky diffuse irradiance is multiplied with a factor (empirical though) to convert the diffuse downwards irradiance to radiance upwards in the direction of the satellite. For calculation of the reflectivity, the ESRA clear sky model is used to calculate the transmissivity downwards to the ground and clouds and upwards to the satellite. An expression for the reflectivity of the thickest clouds is based on measurements from the Nimbus-7 satellite (Taylor & Stowe 1984). No external physical model is used for the ground reflectivity, but rather second lowest value of the reflectivity of a time series for a given pixel. The extreme minimum is avoided because it can be due to artefacts in the processing of the satellite image. The relation between the cloud index and the clear sky index is the same as in the Heliosat-1 version (Equation 6). Thus, despite a physical calculation of the reflectivity, an empirical relation is still used to calculate the global irradiance from the cloud index. 4 2.4 Heliosat-3 (objectives of this thesis) This thesis is a part of the EU-project "Heliosat-3", with the objective of further development of the Heliosat-algorithm to take advantage of the new generation of the Meteosat satellites. While the first seven Meteosat satellites (1977 until present) had only three spectral channels, the next generation (Meteosat Second Generation, MSG) has 12 spectral channels. When the project started in May 2001 the objective was to create an algorithm which in principle consisted of two steps: 1. The new spectral channels of the MSG-satellites should be used to acquire a thorough description of the atmospheric state for any pixel of any image: - clouds (optical depth, coverage, height, phase (water/rain), effective droplet size) - aerosols (type, single scattering albedo, optical depth) - water vapour amount - ozone amount 2. Given a description of the atmospheric state and the solar elevation, the global irradiance and other spectral and angular components should be calculated, preferentially with an advanced Radiative Transfer Model. For the retrieval of the cloud properties, the scheme APOLLO (Saunders et al. 1988, Saunders 1988, Gesell 1989, Kriebel et al. 1989 and Kriebel et al. 2003) is adapted to the MSG satellites. APOLLO was originally developed for the AVHRR sensor of the NOAA satellites, but adaptation to MSG was performed within the project by the German Remote Sensing Data Center (Deutsche Zentrum für Luft und Raumfahrt, DLR), one of partners of the Heliosat-3 consortium. However, the launch of the first of the MSG-satellites was delayed from October 2000 until 28 August 2002. Furthermore, a power supply switched off unexpectedly in October 2002, resulting in even more delay of the operation. Following was a period of commissioning and validation by Eumetsat, and subsequently implementation of the APOLLO algorithm by DLR. Consequently the APOLLO derived cloud parameters were not available for development of a new scheme until after the official end of the project in May 2004. However, the EU extended the project until February 2005, but within the remaining time no improvement of the Heliosat algorithm was achieved by integrating the cloud products in the scheme. Anticipating the delay of data from MSG, the objective within the project was modified to improve the Heliosat-1 scheme based on the calculation of a cloud index. The Heliosat-3 project also deals with the calculation of other solar radiance component like splitting into direct and diffuse radiation, and splitting into spectral components like Ultraviolet, Photosynthetically Active Radiation, Solar Cell Response and Luminance/Illuminance, but this thesis is focusing mainly on the calculation of the horizontal global irradiance. The main objectives of this thesis are then: To develop a clear sky model which can use the operationally retrieved values of aerosols, water vapour and ozone as input. By replacing the simple empirical models used earlier with a numerical Radiative Transfer Model it will also be possible to have spectral output which will ease the subsequent calculation of spectral radiative parameters. To improve the calculation of the cloud index with respect to the following points: - It should be based more on general physical principles and less on tuned empirical parameters. This will ensure that the scheme will also work well for conditions different from those where the tuning has taken place. - It should be as fast and easy to implement as possible for use in an operational scheme. To assess the uncertainty related to sub-pixel variability of the cloud properties. A 3D 5 Radiative Transfer Model will be used to give advice on the ideal size of a pixel and possibly also to suggest modifications to the scheme. In general to improve the accuracy of the calculation of global irradiance, both on a short term but also on longer terms by building the algorithm on physical principles. The algorithm will then be easier to interpret/understand for further development. 3 The concepts of the Heliosat algorithm Like radiation itself, the Heliosat algorithm has a dualistic nature. It can be interpreted as a physcal algorithm, based on the equation for conservation of energy. However, it can also be interpreted as a very simple empirical algorithm, which seems to work well mainly due to statistical cancelling of errors. A very simple - perhaps naive - interpretation of the Heliosat algorithm is given in section 3.1, and in section 3.2 a more strict physical interpretation follows. Section 3.3 is an analysis of how sensitive the output of Heliosat is to the choice of the parameters ρclear and ρcloud (Equation 2). 3.1 An empirical approach A simple interpretation of the Heliosat algorithm is the following: 1. The reflectivity of a Meteosat pixel is calculated by normalising the raw digital counts with incoming radiation at the top of the atmosphere. 2. From a time series the "typical lowest and highest reflectivities" are identified. 3. When the reflectivity is equal to the lowest value the irradiance at ground equals the output of an empirical clear sky model. 4. When the reflectivity is equal to the highest value the irradiance at ground is estimated to be zero. 5. For intermediate reflectivity the irradiance is linearly interpolated between zero and the clear sky value. Some corrections are then added to these simple principles: The scattered radiation from air molecules is subtracted from the normalised digital counts with an empirical formula. This makes the reflectivity more lambertian, at least for the clear sky case, and hence it is easier to compare reflectivities measured under different sunground-satellite geometries. Empirical experience tells us that even under the thickest cloud cover it is never completely dark. When the reflectivity is close to the "typical highest reflectivity" the global irradiance is therefore increased to ca 5-10% of the clear sky value. It is found that by averaging the reflectivity over 3x5 pixels the accuracy of the algorithm is higher when compared to ground measurements. 3.2 A physical approach Conservation of energy implies that solar radiation reaching the top of the atmosphere can be either: 1. reflected to space, 2. absorbed in the atmosphere or 3. absorbed in the ground. 6 In the general case this can be expressed as: I =RG 1− A (7) where I is the incoming irradiance at the top of the atmosphere R is the irradiance reflected to space G is the irradiance reaching the ground (global irradiance) α is the ground albedo A is the radiation absorbed in the atmosphere In the case of no clouds or overcast we have, respectively: I = R clear G clear 1− Aclear (8) I =R cloud G cloud 1− A cloud (9) and For these cases 'clear' and 'cloudy' atmospheres have to be defined for a reference. By neglecting the atmospheric backscatter correction used in Heliosat-1 and Heliosat-2 (Equation 4), and assuming isotropic reflection in all cases, the calculation of the cloud index (Equation 2) is equivalent with: n= R− Rclear R cloud −R clear (10) where radiances have been replaced by irradiances. Solving for these irradiances in Equations 7-9 and inserting into Equation 10 gives: n= 1− Gclear −G Aclear− A 1−G clear −G cloud Aclear− Acloud (11) Solving again for the global radiation G in the general case and introducing the clear sky index k gives: k≡ G cloud n A cloud− A clear Aclear− A G =1−n n G clear G clear 1− G clear (12) By neglecting the variation of atmospheric absorptance (A=Aclear=Acloud) this reduces to: k =1−n G cloud n G clear (13) A typical value for the ratio Gcloud/Gclear is 0.1, but it depends on the solar elevation and the choice of the reference cloud. Equation 13 is not identical to the empirical relation which has been tuned for best performance of the actual Heliosat algorithm (Equation 6), but there are also several differences between the "real world" and this idealised version: Reflectivities are not generally lambertian. 7 A correction for the backscattered radiation from air molecules is used in Heliosat. The absorptance in the atmosphere varies with solar elevation, cloudiness and turbidity. The Meteosat HRV sensor does not measure broadband radiances but is limited to 0.45-1.0 micrometres. Heliosat is tuned to give best correlation between estimated global irradiance averaged over an area (satellite pixel) at a single point in time, and measured global irradiance at ground at a single spot but averaged in time. The theoretical model considers an atmosphere which is homogeneous in the horizontal direction. In reality there are large variations, especially of cloud properties. Thus the empirical relations implicitly account for effects like reduced irradiance due to shadows from nearby clouds, and enhanced radiation due to scattered radiation from broken clouds. The broadband ground albedo α will change slightly when the spectral variation of the incoming irradiance is changing due to varying atmospheric conditions and solar elevation. 3.3 A sensitivity analysis Whether Heliosat is interpreted physically or empirically, the definitions of the reference reflectivities for clear and overcast conditions (ρclear and ρcloud, respectively) are vital. In the various versions of the Heliosat algorithm these parameters are calculated in different ways: In the original version of the Heliosat algorithm (Cano et al. 1986) ρclear and ρcloud were defined as the "most typical values of the reflectivity for clear and overcast conditions". These values were calculated from a time series of the reflectivities with a histogram technique to find the most frequent high and low values from the typical bi-modal distribution. In Heliosat-1 the parameter ρcloud was set to the constant of 160 normalised digital counts for all pixels (Hammer 2000). This value was chosen as the 96 percentile of a time series of all the reflectivity values for the whole field of view of Meteosat. In Heliosat-2 an external empirical model was used for ρcloud since calibrated radiances were used as input instead of uncalibrated digital counts, whereas the parameter ρclear was taken as the second lowest value of a time series. For consistency with Equation 6, a stringent definition of ρclear should be "the reflectivity for which the global irradiance at ground is equal to the clear sky model". Similarly the definition of ρcloud should be "the reflectivity for which the global irradiance at ground is equal to 6.7% of the clear sky model (Equation 6 evaluated at n=1)" The global irradiance calculated with Heliosat is given by Equation 5, which can be written as: G=Gclear k n (14) It is obvious from this equation that the error of the estimated global irradiance is proportional to the error of the clear sky model. Thus an accurate clear sky model is vital. The sensitivity of the estimated global irradiance to the parameter ρclear can be found by differentiating Equation 14: dG dk ∂ n =G clear dn ∂ clear d clear The relative sensitivity is given by: 8 (15) 1 dG 1 dk ∂n = G d clear k n dn ∂ clear (16) Assuming first a simple relationship k = 1 - n, Equations 15 and 16 become, respectively: cloud − dG =G clear d clear cloud −clear 2 (17) 1 dG 1 = G d clear cloud −clear (18) and For the parameter ρcloud the corresponding absolute and relative sensitivities are, respectively: −clear dG =G clear 2 d cloud cloud − clear (19) −clear 1 dG = G d cloud cloud −clear cloud − (20) and Figure 1 shows the sensitivities from Equations 17, 18, 19 and 20 plotted versus the cloud index for typical values of ρclear and ρcloud of 0.2 and 0.8 respectively. On the upper part of the figure one can see that the estimated global irradiance is most sensitive to ρclear for clear cases and to ρcloud for cloudy cases. The magnitude is, however, equal: a 0.01 too high value of ρclear will give an overestimation of global irradiance by almost 2 percent of the clear sky values for clear cases, and a 0.01 too high value of ρcloud will give the same overestimation of global irradiance for the cloudy cases. Hence if one assumes equally many clear and cloudy cases, the average bias of the Heliosat algorithm is equally sensitive to both parameters. However, since the irradiance is lower for the cloudy cases, the relative bias for the cloudy cases is much higher, as seen on the lower part of Figure 1. 9 Figure 1: Sensitivity to the parameters ρclear and ρcloud of the global irradiance estimated with the Heliosat-algorithm. Values of 0.2 and 0.8 are used for ρclear and ρcloud respectively. The upper figure shows the change, in units of the clear sky irradiance Gclear, resulting from a unit change of the parameters ρclear and ρcloud. This is calculated with equations 17 and 19, respectively. The lower figure shows the same, but the unit of change is the fraction of the actual estimated irradiance. This is calculated with equations 18 and 20. Figure 2 shows the sensitivities using the relation between the clear sky index and the cloud index which is used in Heliosat-1 and Heliosat-2 (Equation 6). The relative sensitivity to the parameter ρcloud is now not singular for a cloud index of one, as it is for the simple relation k=1-n. However, the lower part of the figure shows that a value of ρcloud which is too large by only 0.01 will lead to an overestimation of the global irradiance by almost 10 percent for the cloudy cases. 10 Figure 2: Same as Figure 1, but the relationship between the clear sky index and the cloud index is given by equation 6 from Fontoynont et al. (1998). The new equations corresponding to equations 17, 18, 19 and 20 are not shown in the text. 11 4 Summary of results 4.1 Paper I The first paper is mainly a validation of the Satel-Light version of the Heliosat-algorithm (section 2.2). The objective of this work was to identify how well the algorithm is working for various situations and from that find out how it can be improved. To investigate the influence of the sunground-satellite geometry, the performance of the algorithm was analysed for variations of the three angles shown on Figure 3. For overcast situations Heliosat gave too high solar irradiance for low sun and too low irradiance for high sun. For clear situations there were no such biases. In Heliosat there is an implicit assumption that the variation of solar irradiance with solar elevation is similar for clear and cloudy cases. However, numerical simulations with a radiative transfer model showed that irradiance at ground is decreasing faster with solar zenith angle for overcast conditions than for clear conditions. A semi-empirical correction for this was suggested. The deviation between satellite estimates and ground measurements also depends on the solar azimuth angle (or time of day), but much of this variation can be explained by the correlation between the azimuth angle and the solar zenith angle. However, for Bergen it was found that while the modelled global irradiances were symmetric around noon, the measurements were higher in afternoon than in the morning. This asymmetry was only found for intermediate cloudiness, thus suggesting an explanation: even though the average cloudiness did not change with time of day, the scattered clouds were more likely to be positioned over land (to the east) than over sea (to the west). Thus there is greater chance that a given cloud fraction will make shadows at the pyranometer in the morning than in the afternoon. Heliosat relies on the hypothesis that clouds are randomly placed within a pixel, but this "frozen turbulence hypothesis" seems to be violated for Bergen. Empirical corrections for this phenomenon can easily be implemented for a particular site like Bergen, but a general correction is impossible without knowledge of the local conditions for each pixel. The bias is also seen to depend strongly on the co-scattering angle ψ (Figure 3), but again most of this is seen to be due to the correlation with solar zenith angle. When this correlation is corrected for, the opposition effect is seen: when the sun and the satellite are in the same direction (ψ is close to zero) less shadow is seen, and hence the reflectivity is higher. This gives a higher cloud index, and thus the global irradiance is underestimated. The performance of Heliosat was also analysed in light of the total cloud cover and the height of the base of the lowest clouds, as estimated from human observers at ground. For Bergen it was seen that for overcast situations the observed irradiance was smaller than the modelled irradiance. A possible explanation for this is that the clouds in Bergen are very thick, something which also is found by Leontieva et al. (1994). A correction for this can be possible with the second generation of Meteosat satellites (MSG), from which cloud optical thickness can be retrieved by use of more spectral channels. It was also found that Heliosat overestimated global irradiance when the observed height of the cloud bases was very low. Numerical simulations with a radiative transfer model suggest an explanation: by increasing the height of the clouds, the irradiance reaching ground is constant, while the increased reflection is perfectly matched by a decrease of radiation absorbed in the atmosphere. With the MSG satellites cloud (top) height will also become available, and hence a correction for the influence of the cloud height can be implemented. 12 Figure 3: The three angles which are used in this thesis to describe the sun-ground-satellite configuration: solar zenith angle (θ), satellite zenith angle (φ) and co-scattering angle (ψ). 4.2 Paper II One of the objectives of the Heliosat-3 project was to include a Radiative Transfer Model (RTM) directly into the scheme. The advantage of using an RTM is that it can take more detailed input data than the simple models used earlier, and that it can produce spectral output in addition to integrated broadband irradiances. However, it will be unrealistic to run an RTM for each pixel in each image. With approximately 2.5 million pixels to be processed for each image, and an RTM runtime of typically 5 seconds, it would take approximately 3500 hours to process one image. Since MSG produces one image every 15 minutes this approach can not be used in an operational scheme, even with the fastest computers available. Paper II introduces a workaround solution to this problem: an RTM will be used for the clear sky calculations only, and with a spatial resolution of 100 times 100 kilometres. This reduces the runtime to a manageable length, while keeping adequate resolution to describe the spatial variation of atmospheric turbidity. The full resolution will only be used for calculating cloud parameters, which will then be combined with the clear sky calculations with a simple empirical relationship to estimate the radiation at ground. The cloud parameters can then be the traditional cloud index, but also the more physical cloud parameters retrieved with the APOLLO scheme. The paper demonstrates that it is sufficient with two model runs per pixel per day, and that the diurnal variation of the clear sky irradiance can be parameterised with the output from the two model runs. Paper II was mainly written by the first author, Richard Müller. My contribution was discussion and development of the concept of how to reduce the necessary number of runs with an RTM, during a stay in Oldenburg, during project meetings and by email. I also wrote section 4.3 of the paper, a validation of the scheme using daily values of ozone and water vapour and 13 climatological values of aerosol type and optical depth. 4.3 Paper III This paper presents two modifications to the algorithm which is used to calculate the cloud index. The first modification concerns the backscatter correction in the Heliosat scheme (Equation 4). In the calculation of the cloud index the part of the satellite signal that is scattered from the atmospheric molecules is subtracted. This correction was introduced by Beyer et al. (1996), and an empirical expression was developed. By assuming that the reflected radiance from the sea surface is negligible to the contribution scattered from the air molecules, Beyer et al. fitted an expression to the count values of Meteosat HRV images for a number of cloud free pixels over sea in Western Europe for August 1993. In Hammer (2000) this expression was modified by fitting to a larger database. Paper III of this thesis replaces these empirical expressions with an analytical expression. This is possible by assuming a plane-parallel atmosphere and that multiple scattering is negligible for the wavelengths of the satellite sensor. The expression is derived for monochromatic radiation, but it is demonstrated that by use of an "equivalent wavelength" it is not necessary to perform integration over the spectral region of the sensor. While the empirical expressions used earlier are tuned for a particular satellite sensor, the new analytical expression can easily be adapted to other sensors. Furthermore, while the empirical equations are fitted to certain sun-ground-satellite configurations (Figure 3), the new expression is valid for all angles for which the plane-parallel approximation is reasonable. The second modification concerns the calculation of the reflectivity of the ground for each pixel. In earlier versions of Heliosat this parameter (ρclear) was determined for each month and each slot. This assures that the sun-ground-satellite geometry is fairly constant, and therefore avoids partly the problem of non-lambertian reflection. However, this is a time consuming part of the algorithm, and besides it can be difficult to determine ρclear for months and slots with few clear situations. In Paper III it is demonstrated that the lower bound of reflectivity can be parameterised with the coscattering angle (ψ on Figure 3). Thus the ground albedo for each pixel can be calculated once and for all, and the algorithm will then be significantly faster and simpler to implement. Global irradiances are calculated using both the modified cloud index from this paper and the traditional cloud index from the Heliosat-1 version (Section 2.2). Two different clear sky models are also used for the calculations: the model from Heliosat-1 and the new SOLIS model developed in Paper II, both using climatological input of turbidity data. When compared to ground measurements at five European stations, it is found that the accuracy is quite similar for all of the four combinations of the two cloud indices and the two clear sky models. However, it is found that the new cloud index is on the average higher than the traditional one, particularly for low values as it is less frequently negative, and that the SOLIS model gives generally lower clear sky values than the traditional clear sky model. Consequently the new cloud index gives best results when combined with the old clear sky model, and vice versa. It is therefore important to be aware of this when combining a cloud index with a clear sky model in the Heliosat scheme. 4.4 Paper IV Paper IV gives a validation of the database HelioClim-2, which provides global irradiance for the solar energy and daylight community on the internet. The HelioClim-2 data is calculated with the Heliosat-2 algorithm (section 2.3) using Meteosat images which are sampled in both time and space as input. The sampling is done to be able to efficiently process long time series of data for the whole field of view of Meteosat for climatological purposes. In this paper five European ground stations are used for the comparison. 14 For daily values of global irradiance HelioClim-2 performs much better than the previous version of the database (HelioClim-1) which used a lower sampling frequency in both time and space. However, HelioClim-2 gives a larger Root Mean Square Deviation than Heliosat-1 (section 2.2). Most of this difference is probably due to the sampling of the satellite data, but some of it can also be due to differences between the Heliosat-1 and Heliosat-2 algorithms. HelioClim-2 overestimates the global irradiance for all five ground stations, and for the station Bergen the bias is as large as 21%. Since there are also more cloudy cases in Bergen than for the other sites, it is suggested that Heliosat-2 overestimates the global irradiance for the cloudy cases. It is shown that this overestimation can be due to a too high value for the cloud reflectivity. This paper is a draft only. The second author, Lucien Wald, will later conclude the paper with a discussion of the effects of interpolating and sampling the satellite data which are input to HelioClim-2. Besides, HelioClim-2 will be reprocessed with a lower value of the cloud reflectivity, as suggested in the draft, to see if the bias will be smaller. 4.5 Paper V The greatest challenge of the Heliosat algorithm is that cloud properties are varying on a scale which is often smaller than the size of the satellite pixels. This sub-pixel variability cannot be accounted for directly, since detailed information is missing. Nevertheless it is of interest to have an estimate of the variability of the reflected radiance for different sizes of the pixels. In this paper a 3dimensional radiative transfer model, SHDOM, is used to simulate the reflectance towards a satellite from two different cloud fields. One is a rather homogeneous stratocumulus field, but with typical 'bumps' on the top, and the second is a broken cumulus field with only scattered clouds. With the sun and 'satellite' in fixed positions, the cloud fields are rotated in the azimuth direction. It is found that by observing the full stratocumulus field of 3520 metres, there is no variation of the reflectance. For the cumulus cloud field the variability is approximately 10-20 percent, even though the size of this field is larger; 6700 metres. Looking at smaller subsets of the cumulus cloud field, the variability is extreme, while for the stratocumulus field it is more moderate. In Paper V the distribution of reflected radiance from the stratocumulus field in different directions is also compared with reflectances measured with Meteosat. When plotted versus the co-scattering angle (ψ on Figure 3) it is found that the reflectance of the thickest clouds measured with Meteosat is rather homogeneous, but with a minimum around ψ = 60 degrees. The simulated reflectances depend on ψ in a qualitatively similar way, but the variation is much larger: the reflectivity for ψ = 60 degrees is only 50 percent of the reflectivity for ψ close to zero (i.e. when the sun and the satellite are in the same direction, as seen from the cloud field). One of the objectives of this study was to find a parameterization of ρcloud, the albedo of the thickest cloud used in Heliosat. However, it was found from the Meteosat reflectivities that the thickest clouds were close to lambertian reflectors while the angular distribution of radiances was more variable for low and intermediate cloudiness. Hence, it should make more sense to parameterise the cloud index - clear sky index relationship with ψ than parameterising the reflectivity of the thickest clouds. 5 Concluding remarks Heliosat is an algorithm for estimating solar radiation at ground from images taken in the visible range by geostationary satellites. This algorithm consists in principle of two parts: 1) a model to describe the radiation at ground level when there are no clouds, and 2) a method to combine the output of this clear sky model with a measure of the cloud cover to estimate the solar radiation in the general case when there are clouds. Paper II of this thesis shows a new clear sky model scheme, SOLIS, which is based on a numerical 15 Radiative Transfer Model, replacing the simple empirical models used earlier. Firstly, this permits the use of more detailed information about aerosols (dust), water vapour and ozone as input, parameters that in the near future will be retrieved operationally using the new generation of the Meteosat satellites. Secondly, the model also provides spectral output which will be useful for deriving specific solar radiation parameters like Ultraviolet radiation, Photosynthetically Active Radiation, Illuminance (visible light) and Solar Cell response. The SOLIS model is seen (in papers II, III and IV) to give reasonable results using climatological input parameters. However, the potential of the scheme can only be released in the near future when operationally retrieved atmospheric parameters are used as input. Paper III suggests two modifications to the traditional algorithm for calculation of the 'cloud index'. First, a parmeterisation of the effect of non-lambertian reflection from ground makes the algorithm significantly faster and easier to implement. Second, an analytical correction for the radiation scattered from the atmospheric molecules towards the satellite is derived, replacing an empirical expression in the original algorithm. This makes the algorithm more general, and also easier to interpret and develop further. A validation versus ground measurements shows that the accuracy using the new cloud index is similar to using the old one. However, it is also demonstrated how biases which are due to the cloud index and the clear sky model can cancel or add up, and thus it is important that these two separate components of the algorithm are "kept in balance" with each other. Papers I and IV are validations of several different versions of the Heliosat algorithm. This has given a basis for understanding the algorithm and several corrections and modifications are suggested. Paper V is an assessment of the influence of sub-pixel variability of cloud properties on the reflectance. 6 Bibliography Beyer H.G.; Costanzo C.; Heinemann D;, 1996: Modifications of the heliosat procedure for irradiance estimates from satellite images, Solar Energy, 56, 207-212. Cano D; Monget JM;, Albuisson M; Guillard H; Regas N; Wald L;, 1986: A method for the determination of the global solar radiation from meteorological satellite data, Solar Energy, 37, 3139. Dumortier, D, 1998: The Satel-Light model of turbidity variations in Europe. Report for the 6th Satel-Light meeting in Freiburg, Germany, September 1998 Dumortier, D, 1995: Modelling global and diffuse horizontal irradiances under cloudless skies with different turbidities. Technical report for the Daylight II project, JOU2-CT92-0144 Fontoynont, M; Dumortier, D; Heinemann, D; Hammer, A; Olseth, J A; Skartveit; Ineichen, P; Reise, C; Page, J; Roche, J; Beyer, H; Wald, L, 1998: Satel-Light: A www server which provides high quality daylight and solar radiation data for western and central Europe. Proc. 9th conference on satellite meteorology and oceanography in Paris, 25-28 May 1998, pp. 434-437 Geiger M.; Diabaté L.; Ménard L.; Wald L., 2002: A Web service for controlling the quality of global solar irradiation, Solar Energy, 73 (6), 475-480. Gesell, G., 1989: An Algorithm for Snow and Ice Detection Using AVHRR Data: An Extension to the APOLLO Software Package, International Journal of Remote Sensing, 10 (4-5), 897-905. Grüter, W.; Guillard, H.; Möser, W.; Monget, J-M.; Palz, W.; Raschke, E.; Reinhardt, R.E.; Schwarzmann, P. and Wald, L., 1986: Solar Radiation Data from Satellite Images. Solar Energy R&D in the European Community, Series F, Volume 4, D. Reidel Publishing Company, p. 100 Hammer, A, 2000: Anwendungsspezifische Solarstrahlungsinformationen aus Meteosat-daten, PhD thesis, University of Oldenburg. 16 Kriebel K. T., Gesell G., Kästner M., Mannstein H, 2003: The cloud analysis tool APOLLO: Improvements and Validation, Int. J. Rem. Sens, 24, 2389-2408. Kriebel, K.T.; Saunders, R.W.; Gesell, G., 1989: Properties of Clouds Derived from Fully Cloudy AVHRR Pixels, Beiträge zur Physik der Atmosphäre, 62 (3), 165-171. Lefèvre M.; Bauer, O.; Iehle, A.; Wald, L., 2000: An automatic method for the calibration of time-series of Meteosat images, International Journal of Remote Sensing, 21 (5), 1025-1045. Leontieva, E; Stamnes, K; Olseth, J A;, 1994: Cloud optical properties at Bergen (Norway) based on the analysis of long-term solar irradiance records, Theoretical and Applied Climatology, 50, 7382. Page, J, 1996: Algorithms for the Satel-Light programme. Technical report for the Satel-Light programme Rigollier C; Bauer O; Wald L;, 2000: On the clear sky model of the 4th European Solar Radiation Atlas with respect to the Heliosat method, Solar Energy, 68(1), 33-48. Rigollier C.; Lefèvre M.; Blanc Ph.; Wald L., 2002: The operational calibration of images taken in the visible channel of the Meteosat-series of satellites., Journal of Atmospheric and Oceanic Technology, 19 (9), 1285-1293. Rigollier, C; Lefèvre, M; Wald, L, 2004: The method Heliosat-2 for deriving shortwave solar radiation from satellite images, Solar Energy, 77, 159-169. Saunders, R.W, 1988: Cloud top temperature/height: A high resolution imagery product from AVHRR data, Meteorological Magazine, 117, 211-221. Saunders, R.W. and K.T. Kriebel, 1988: An improved method for detecting clear sky and cloudy radiances from AVHRR data, International Journal of Remote Sensing, 9, 123-150. Scharmer, K;, 1994: Towards a new atlas of solar radiation in Europe, Solar Energy, 15, 81-87. Taylor, V.R.; Stowe, L.L., 1984: Atlas of reflectance patterns for uniform Earth and cloud surfaces (Nimbus 7 ERB - 61 days). NOAA Technical report NESDIS Wald, L.; Wald, J-L. and Moussu, G., 1992: A technical note on a low-cost high-quality system for the axquisition and digital processing of images of WEFAX type provided by meteorological geostationary satellites, International Journal of Remote Sensing, 13, 911-916. Worldwatch Institute, 2003: State of the World 2003, New York: W.W. Norton & Company. Zelenka A.; Perez R.; Seals R.; Renne D., 1999: Effective accuracy of satellite-derived hourly irradiances, Theor. Appl. Climatol., 62, 199-207. 7 Part II - The papers I-V 17 Paper I Dagestad, K-F. (2004) Mean bias deviation of the Heliosat algorithm for varying cloud properties and sun-ground-satellite geometry Theoretical and Applied Climatology, 79 215–224 Theor. Appl. Climatol. 79, 215–224 (2004) DOI 10.1007/s00704-004-0072-5 Department of Geophysics, University of Bergen, Bergen, Norway Mean bias deviation of the Heliosat algorithm for varying cloud properties and sun-ground-satellite geometry K.-F. Dagestad With 10 Figures Received October 26, 2003; revised April 5, 2004; accepted May 20, 2004 Published online September 29, 2004 # Springer-Verlag 2004 Summary Reflectances measured by the ‘visible’ channel of Meteosat are converted to global radiation at the ground by the Heliosat-algorithm. This algorithm is based on the inverse relationship between reflectance and transmittance. This relationship is, however, complicated by two factors: 1) the absorptance of the atmosphere is varying, and 2) the reflectivity is different in different directions. These two factors are again depending on the state of the atmosphere (mainly clouds) and the sun-ground-satellite geometry. The performance of the Heliosat-method is here tested against ground data from Bergen (Norway), Geneva (Switzerland) and Lyon (France), and analysed in light of cloud properties as observed from ground, and different sun-ground-satellite geometries. Situations where modelled and measured radiation differ are identified, and possible causes are suggested. Some of the suggestions are supported by calculations with a radiative transfer model. 1. Introduction Geostationary satellites have become a valuable tool for estimation of global radiation at ground level. Although ground measurements often will be the most accurate for a fixed position, the network of measuring sites is sparse, especially over sea. The spatial coverage is substantially improved by a geostationary satellite, which covers almost half of the earth, with typical pixel sizes of 5 5 km2. The satellite vs. ground truth Root Mean Square Deviation (RMSD) for hourly global irradiance is typically 20–25%, which is comparable to the accuracy obtained by interpolation from ground stations some 20–30 km apart (Zelenka et al., 1999). Still there is some potential for improvement; through newer satellites (more channels, smaller pixels, higher time resolution) and refinement of the conversion algorithms (e.g. Heliosat). With the Meteosat Second Generation (MSG) satellites, more information about the atmospheric state (including clouds) will be available. This information can improve the scheme, if the effect of the various parameters can be properly accounted for. For this paper the performance of the Heliosat method is analysed with respect to variations of the Mean Bias Deviation (MBD) due to different cloud properties or sun-ground-satellite geometries. Heliosat-derived hourly global irradiances are compared to ground measurements in Bergen, Norway (60.4 N, 5.3 E), Geneva, Switzerland (46.2 N, 6.1 E) and Lyon, France (45.8 N, 4.9 E) for the years 1996 and 1997. The performance of the Heliosat method was tested against seven Norwegian and five other European stations by Olseth and Skartveit (2001). Tests of satellite based estimates against ground truth data have also been done by Ineichen and Perez (1999). These papers did not, however, use infor- 216 K.-F. Dagestad * Fig. 1. Three of the angles used in this paper: solar zenith angle , satellite zenith angle and co-scattering angle . Not shown on this figure is the solar azimuth angle, which relates the solar position to the north-direction mation about observed ground truth cloud properties and the sun-ground-satellite angles, except for the solar elevation. Three angles affect the signal received by Meteosat: 1) the solar zenith angle , 2) the satellite zenith angle , and 3) the ‘co-scattering angle’ , between the direction towards Meteosat and the sun as seen from ground. These angles are shown on Fig. 1. Of these, only the solar zenith angle affects the global radiation measured at ground. The signal received by the Meteosat VISsensor (0.5–0.9 mm) is scattered=reflected from four different scattering agents: * * * Clouds are the main reflectors of solar radiation in the earth-atmosphere system. They absorb moderately, but the scattering is very strong. Angular distribution of scattering from individual cloud droplets is well known from theory (Mie), but ice crystals, multiple scatter and macroscopic cloud structure make the picture more complex. Aerosols, like cloud droplets, are also scattering strongly forward. They are, however, more evenly distributed, and exist in smaller amounts. The collective scattering from an aerosol layer is therefore closer to the angular distribution given for individual particles by the Mie-theory. The third scattering agent is the air molecules. Because of their smaller sizes they scatter solar radiation according to the theory of Rayleigh; symmetric in the forward- and backward directions, with smaller amounts to the sides. The satellite signal also contains radiation reflected from the ground. For terrestrial surfaces the reflection is nearly isotropic (Lambertian). However, for rough surfaces, there will be strong backward reflection, since an observer (satellite) in this position will see less shadows (opposition effect). For ocean the reflection can be strong in the forward direction if the sun is low (specular reflection). Since each of these four scattering agents has different angular distributions, it is difficult to find a general relationship between the radiance towards a satellite and the top of atmosphere reflection. However, some semi-empirical corrections exist in the Heliosat-scheme (next section). Some of the solar radiation is also absorbed in the atmosphere, typically 10–15% for high sun and up to 20–30% for low sun. The main absorbers are water vapour, aerosols, and ozone. Quantitatively water vapour is most important, but since the most intense water vapour absorption bands are quickly saturated, the absorptivity due to water vapour is normally not very variable. Aerosols are more variable, however, both in amount and type, and therefore changing aerosol conditions may have large impacts on the global radiation. Absorption by ozone is generally small for global solar irradiances, but may be important for a satellite received signal in case of dense cloud cover and low sun, when the radiation makes a long path through the ozone layer, both downwards and upwards. The total absorptivity is sensitive to the solar zenith angle, which determines the path length of the photons. A change of the atmospherical absorptivity disturbs the inverse relationship between reflectance and transmittance. Since it is not directly measured it has to be estimated by an empirical formula. 2. The Heliosat method The Heliosat method was first proposed by Cano et al. (1986). It has been modified several times, and is still under development. The version Mean bias deviation of the Heliosat algorithm presented here is the version used by the SatelLight project (www.satel-light.com; Iehle et al., 1997; Hammer et al., 1998, 2003). The method consists of two clearly separated steps: 1. First a ‘cloud index’ is determined from time series of pixel counts from the VIS-channel of Meteosat (0.5–0.9 mm). The cloud index is a measure of the relative reflectivity of the clouds (in the direction of Meteosat). 2. The cloud index is combined with an empirical clear sky model to give the global irradiance at ground. The details are as follows: 2.1 Finding the cloud index The relative reflectivity of a pixel is defined by: C C0 Catm ð1Þ " cos where " is the correction for the sun-earth distance, C is the raw pixel count of Meteosat (an integer between 0 and 255), and C0 is an instrument offset used to adjust the null point of the sensor. Catm is the contribution scattered from the cloud free atmosphere. This part is subtracted because of its highly anisotropic character. Catm has been estimated with semi-empirical formulas, depending only on the three angles shown on Fig. 1. For the Satel-Light project the following formula was used: Catm ¼ ð1 þ cos 2 217 and no correction are made for e.g. the opposition effect. The cloud index is 1 when ¼ c and zero when ¼ g. Due to the anisotropy of the reflected radiation, n can thus be smaller than 0 or greater than 1. 2.2 Determining the global radiation from the cloud index The Heliosat method is based on an empirical relationship between the cloud index and the clear sky index k defined by: G ð4Þ k Gclear where G is the actual global radiation and Gclear is global solar irradiance estimated by a clear sky model. Input data to this model are the solar zenith angle and climatologial monthly values of the Linke turbidity factor at the relative optical air mass of 2. The direct part of the clear sky model is given by Page (1996), and the diffuse part by Dumortier (1995). For the direct part also the height above sea level is used as an input parameter. The relation between n and k used by the Satel-Light project is: k ¼ 1:2 for n 0:2 k ¼1n for 0:2n0:8 2 ð5Þ for 0:8n1:1 k ¼ 3155nþ25n 15 k ¼ 0:05 for n 1:1 ———————————————————— 2 Þð0:55 þ 25:2 cos 38:3 cos þ 17:7 cos 3 Þ ð2Þ cos 0:78 ———————————————————— The cloud index is then defined by: g n c g ð3Þ where is the relative reflectivity of the actual pixel, and g and c is the relative reflectivity of the same pixel with no clouds and overcast conditions, respectively. g is taken as the most frequent value for each pixel, and different values are used for each month. c is the relative reflectivity of a very dense cloud cover; a constant value (of 160) was chosen for all pixels, taken as the 96 percentile point for all pixels from a time series. This means that both clouds and ground are treated as Lambertian in Heliosat, This formulation ensures that G ¼ Gclear when n ¼ 0 and not less than 5% of Gclear even under the thickest clouds. The role of the clear sky model is actually to estimate the absorption in the atmosphere, which is not measured directly. A k ¼ 1 n relation is equivalent to assuming that the absorptivity does not change with cloudiness. 3. Description of data 3.1 Observed global irradiances For all three sites global radiation is measured with Kipp & Zonen pyranometers; CM6 in Lyon and 218 K.-F. Dagestad CM10=CM11 in Geneva (http:==idmp.entpe.fr) and Bergen (Radiation Yearbooks 1996–1997). For Bergen the measurements are given as hourly values in solar time, where e.g. the values at 12 means measurements between 11:30 and 12:30. For Geneva and Lyon the timing of the measurements was different, but the data were summed and weighted to yield the same time format as for Bergen. To ensure high quality of the data, only hours with the sun entirely above the horizon are used, with the second criteria that the mean solar elevation for the hour is also at least 10 . Data from November to April were not used for any of the sites due to potential snow cover. 3.2 Global irradiances from Heliosat From the Satel-Light webpage (www.satel-light. com), estimates of global irradiances were downloaded for the three sites for the years 1996 and 1997. Data are given each 30 minutes (repetition cycle of Meteosat). Hourly values of the Heliosatestimates were created by weighting with the number of minutes they overlap a given hour. Each satellite measurement was assumed to have a ‘radius of coverage’ of 15 minutes, which means that e.g. the hour 12 (11:30–12:30) is covered by satellite measurements within the interval 11:15–12:45. The pixel size for the ‘visible’ Meteosatchannel is 2.5 2.5 km2 at nadir, with increasing size away from the sub-satellite point. In Central Europe (like Geneva and Lyon) a pixel covers approximately 2.5 km in longitude and 4 km in latitude. For Bergen the corresponding numbers are 2.5 km in longitude and 7 km in latitude. Tests within the Satel-Light project showed that best results were obtained by averaging 5 pixels in longitude and 3 pixels in latitude. So for the data used here Heliosat was applied to pixels of approximately 12 12 km2 for Geneva and Lyon and approximately 12 21 km2 for Bergen. The averaging of pixels makes the covered area more equal to the area of the sky ‘‘seen’’ by an instrument on ground, which is typically 50 50 km2. 3.3 Observed cloud properties The third independent data source is the cloud observations from ground. Every three hours (9, 12. . .UTC) many parameters are observed, of which two are used here; the fractional cloud cover (given in octa) and the height of the base of the lowest clouds. Note that these are only rough estimates by a human observer. For Bergen the cloud observations are taken on the same site as the radiation measurements. In Geneva the station is 5.5 kilometres away and in Lyon it is less than 5 kilometres away. Data are provided by the Norwegian Meteorological Institute, Meteo Suisse and Meteo France respectively. 4. Comparison between Heliosat and ground measurements The hourly global irradiances are normalized with the clear sky model of Satel-Light (Section 3.2), and the resulting clear sky indices are denoted by ksat for the Heliosat derived values and kobs for the ground measured values. The term ‘deviation’ is in this section defined by ksat kobs . In the following sections the Mean Bias Deviation (MBD) is analysed in light of the solar zenith-, solar azimuth- and co-scattering angles and the observed cloud amount and height of the cloud bases. 4.1 Solar zenith angle Figure 2 shows the MBD plotted against solar zenith angle for observed cloud amount of 0 Fig. 2. Mean Bias Deviation (MBD) of the satel-light derived clear sky indices versus the ground observed counterparts ðksat kobs Þ for Bergen, Geneva and Lyon plotted versus the solar zenith angle. Upper part is for ‘clear’ cases, with observed cloud amount of 0 or 1 octa. Lower part is for ‘cloudy’ cases with observed cloud amounts of 4–8 octa Mean bias deviation of the Heliosat algorithm or 1 octa (‘clear’) and 4 octa (‘cloudy’). For Geneva and Lyon the bias is small for the clear cases, except a positive bias for low sun. For the cloudy cases Heliosat gives too small values for high sun (<50 ) and too high values for low sun. The reason for this could be that either the reflectance (and hence cloud index) or the global irradiance at ground (and hence clear sky index) varies differently with with and without clouds. To test these hypotheses, a small case study was performed with the radiative transfer model SBDART (Richiazzi et al., 1998). Reflected top of atmosphere irradiance and global irradiance at ground was simulated for different solar zenith angles for a clear sky case and for a cloudy case with various cloud optical depths. The simulations showed that the dependence of reflected irradiance on is almost completely similar with and without clouds (figure not shown). The global irradiance at ground, however, varies differently with for the clear and cloudy case, as seen on Fig. 3. The values are normalized so that the areas under the curves are equal, so that only the relative variation with is shown. The global irradiance at overcast is higher for high sun and lower for low sun, relatively to the clear case. The ‘crossing’ appears around a solar zenith angle of 35 . This might be the most likely explanation for the variation of the bias shown on Fig. 2. Furthermore, Fig. 3. Global irradiance at ground versus solar zenith angle simulated with SBDART with and without clouds. The values are normalized so that the area under the curves are equal. The clouds used were water clouds with an optical depth of 20 for the wavelength of 0.55 mm 219 it was found that already at an optical depth of 4 the variation of global irradiance with theta did not change with further increase in cloud thickness. One must remember two things, however: the clouds in SBDART are plane-parallel and reflected irradiance was used instead of radiances in a certain direction. A physical explanation for the different solar zenith angle dependence of global irradiance for clear and cloudy skies can be found from Skartveit and Olseth (1996); Their model of diffuse sky irradiance on an inclined plane assumes that at overcast conditions 30% of the horizontal diffuse irradiance is due to collimated radiation from zenith. This ‘zenith brightening’ for overcast conditions appears at all solar zenith angles. Therefore, when the sun is located close to zenith, where the cloud transmissivity is largest, the global irradiance is relatively high. Accordingly the transmissivity decreases more rapidly as the sun is moving away from zenith for overcast than for clear conditions. A simple empirical correction for this bias can be made based on Fig. 2, by replacing ksat by ksat þ F1 ðÞ for overcast situations. However, based on Fig. 3 a more physical correction can be deduced, by multiplying the clear sky index for overcast conditions with a correction describing the effect of the zenith brightening: replacing ksat by ksat F2 ðÞ No attempt will be made to determine such functions here, based on a sparse data set and a simple case study with SBDART. For Bergen the cloudy cases are qualitatively similar to the cloudy cases in Geneva and Lyon, except that the variation of the MBD with is not as strong in Bergen as it is in Geneva and Lyon. For the clear cases in Bergen Heliosat gives too low global irradiance for low sun. Bergen differs from the other two sites by higher latitude and higher satellite zenith angle. The radiance scattered from a long atmospheric column can be comparable to reflection from clouds when the satellite zenith angle and solar zenith angle are very large, and Heliosat may overestimate the cloud amount for these cases. In other words: the correction for backscatter from the atmosphere (Eq. 2) may give too low values when both the sun and satellite are close to the horizon. As a consequence, the cloud index will be too high, and the estimated irradiance too low, as is the case on Fig. 2. 220 K.-F. Dagestad Fig. 4. Mean Bias Deviation (MBD) of the satel-light derived clear sky indices versus the ground observed counterparts ðksat kobs Þ for Bergen, Geneva and Lyon plotted versus the solar azimuth angle 4.2 Solar azimuth angle Figure 4 shows the MBD plotted against solar azimuth angle. This angle is here defined as 0 when the sun is to the north, and increases during the day with a value of 180 when the sun is to the south. For Geneva and Lyon we see a similar shape; Heliosat gives too high values in the morning and afternoon, and too low values around noon ( ¼ 180 ). Most probably this is due to the relationship between the solar azimuth angle and the solar zenith angle; the solar zenith angle is lower when the sun is in the south (mid of the day). For Bergen, however, there is a ‘strange’ diurnal trend of the bias, with Heliosat giving too high estimates in the morning and too low values in the afternoon. It is found, furthermore, that Heliosat-estimates (ksat) are symmetrical about noon (azimuth ¼ 180 ), while observed clear sky indices (kobs) have the asymmetry reflected in Fig. 4. Looking more closely, this asymmetry is not present for small cloud amounts or for fully overcast, but only for intermediate cloud amounts. For observed cloud amount in the range 4–7 Fig. 5 shows that the mean observed clear sky index in Bergen ranges from 0.53 in the morning to 0.68 in the afternoon. Since a similar asymmetry does not exist for totally overcast, the reason is probably the cloud position on the sky, rather than different cloud thickness. Since Bergen is situated on the west coast, most probably at this time of the year (summer=autumn) there are more frequently clouds to the east (over land) than to the west (over sea). Observers at the Meteorological Institute in Bergen confirm that this is also their experience (personal communication), although no quantitative measurements of this phenomena exists. Heliosat relies on the assumption that clouds are randomly placed in space, probably a good assumption, except for orographic clouds or for clouds along land=sea boundaries. To correct for such phenomena, Heliosat needs very small pixels. This again introduces other problems, such as the need to correct for the shadows made by the clouds, and thus generally gives higher RMSD. Some of the variations in Fig. 5 are also due to the simple fact that the mean cloud amount is somewhat higher in the morning than in the afternoon. The mean cloud amount for the dataset on Fig. 5 is ranging from 5.8 to 6.3 with the higher values in the morning. Based on numbers from Olseth and Skartveit (1993) this alone cannot account for the large variation in the clear sky index. 4.3 Co-scattering angle Fig. 5. Mean observed cloud index in Bergen for each hour with observed cloud amounts in the range of 4–7 octa The angle between the direction towards the satellite and the sun, as seen from ground, is here called the ‘co-scattering’ angle. ( on Fig. 1). It is physically meaningful to investigate this angle since the phase functions for scattering by molecules, aerosols and cloud droplets depend on it. Also the amount of shadows seen on ground and clouds varies mainly with . When is zero, no shadows are seen from the satellite, and the reflectivity should reach a peak value (opposition effect). Figure 6 shows the MBD plotted against the co-scattering angle. Here the clear (observed Mean bias deviation of the Heliosat algorithm Fig. 6. Mean Bias Deviation (MBD) of the satel-light derived clear sky indices versus the ground observed counterparts ðksat kobs Þ for Bergen, Geneva and Lyon plotted versus the co-scattering angle. Upper part is for ‘clear’ cases, with observed cloud amount of 0 or 1 octa. Lower part is for ‘cloudy’ cases with observed cloud amounts of 4–8 octa cloud amount 1=8) and cloudy (observed cloud amount 4=8) cases are separated. For Geneva and Lyon the bias is close to zero for the clear cases, except for slightly negative values for <5 and positive values as increases above 70 . For the cloudy cases the picture is similar, except for a ‘strange dip’ at around 30 . This can be explained through the relationship 221 between and , as shown on Fig. 7: For around 30 , the mean value of the solar zenith angle is at a minimum. So the bias shown for the cloudy cases on Fig. 2 also shows up on Fig. 6 via the dependence. When an empirical correction based on Fig. 2 is made; ksat ¼ ksat =400 þ 0:105 for cloud amount larger than 2, the bias for the cloudy cases on Fig. 6 is similar to the clear cases (figure not shown). The only remaining ‘feature’, the negative MBD for <5 , can be explained by the opposition effect: No shadows are seen from Meteosat when the sun is in the same direction, the reflectivity is larger, and hence the cloud index is too high and the estimated global irradiance too low. However, less than 2% of the hours have less than 5 , so the influence on the overall performance of Heliosat is small. Since the ‘dip’ in the MBD at low values of appears both with and without clouds, the opposition effect occurs both on clouds and on the ground. For Bergen most of the variation of the MBD with is also due to the relation between the co-scattering angle and the solar zenith angle (Fig. 7). However, since the MBD varies differently with in Bergen compared to Geneva and Lyon (Fig. 2), it also looks different on Fig. 6. But also in Bergen a clear sign of the opposition effect is seen for less than 5 . 4.4 Cloud cover Figure 8 shows MBD versus total cloud amount observed from ground. For all three sites the MBD is fairly constant, but rises when the cloud amount is 8=8 (Heliosat gives too high global irradiance). For Bergen the increase is larger, and also starts at cloud amount 7=8. Two theories Fig. 7. Correlation between the solar zenith angle and the co-scattering angle. The ordinate is the mean of the solar zenith angle within each ‘bin’ of the co-scattering angle (abscissa). The borders of the bins equal the x-grid Fig. 8. Mean Bias Deviation (MBD) of the satel-light derived clear sky indices versus the ground observed counterparts ðksat kobs Þ for Bergen, Geneva and Lyon plotted versus the total cloud amount given in octa 222 K.-F. Dagestad to explain the overestimation by Heliosat for fully overcast are proposed here: 1. For cloud amount less than 8=8, at least some direct radiation reaches ground. Some of this is then reflected, and some is even reflected back to ground again. Such multiple scattering=reflection might in some cases even augment the global irradiance above the incoming irradiance on the top of the atmosphere (for short-term values). A full cloud cover caps off the direct source, and it becomes darker at ground without the reflectivity increasing proportionally. 2. The thickest clouds most probably also have cloud amounts of 8=8. After ‘the hole is filled’, the clouds might thicken to make it very dark at ground, without the reflectivity rising enough to compensate. Simulations with SBDART (not shown) support this hypothesis: The decrease of radiation reaching ground when cloud thickness increases very much are mainly ‘lost’ due to absorption within the clouds rather than due to reflection to space. It is reasonable that Bergen has the highest increase of MBD for totally overcast, since it is a coastal city placed in the westerlies at 60 N. Leontieva et al. (1994) showed that the clouds in Bergen have high optical thickness. MSG data will make it possible to retrieve estimates of the cloud optical depth. This parameter might be a valuable replacement or addition to the cloud index. For broken clouds, the MBD has a rather constant level for all sites. This level is mainly set by the clear sky model with its climatologically input of the Linke turbidity coefficient. Equation 4 shows the sensitivity of the clear sky model on the performance of Heliosat: any error in the clear sky model gives the same error on the modelled global irradiance. Apparently the Linke turbidity coefficient fits well for Lyon, while it is too low for Geneva and too high for Bergen. In the future, with the use of the new MSG satellites, hopefully more accurate real time input to the clear sky models will be available. 4.5 Height of cloud base of lowest clouds Figure 9 shows the MBD plotted against the height of the base of the lowest clouds as estimated by an observer at ground. The bias is not Fig. 9. The upper part shows the Mean Bias Deviation (MBD) of the satel-light derived clear sky indices versus the ground observed counterparts (ksat kobs) for Bergen, Geneva and Lyon plotted versus the observed (estimated) height of the base of the lowest clouds. The lower part shows the histogram of the number of hours within each ‘bin’, where the borders of the bins equal the x-grid very large for medium and high clouds, but increases substantially for height of the cloud base lower than 600 metres. However, as Fig. 9 also shows, there are few cases with the cloud base below 600 metres; between 12% and 18% for the three sites. Consequently the effect on the overall bias is not very large, but it is still interesting to examine the physics behind it. With Fig. 10. Reflection, absorption and global irradiance simulated with SBDART versus the height of the cloud base. The solar zenith angle is 0 and the clouds have an optical depth of 20 for the wavelength 0.55 mm. For all heights of the cloud base the clouds have a geometrical thickness of 1 km (one layer in the model). All values are normalized with the incoming irradiance at the top of atmosphere for the corresponding solar zenith angle Mean bias deviation of the Heliosat algorithm SBDART the height of the clouds can be changed keeping everything else constant. Figure 10 shows the changes of reflectivity, global irradiance and absorptivity in the atmosphere with changing height of the clouds. The fraction of the incoming irradiance that reaches the ground is more or less unaffected by the cloud height. The reflectivity, however, increases with cloud height, and this is matched by a decrease in the absorptivity. This means that higher clouds ‘shelter’ the solar radiation from the absorbing species in the lower atmosphere. The height of the clouds affect the reflectivity (cloud index) without affecting the global irradiance (clear sky index), hence the relationship given by Eq. 5 is disturbed. Figure 9 shows that the effect is only obvious for cloud heights below 600 metres. This makes sense, since both aerosols and water vapour normally decrease exponentionally with height. The SBDART simulations, however, indicate that there should be an effect also by raising the clouds all the way to the tropopause. While the cloud properties (optical- and geometrical thickness, effective radii and water phase) are kept constant in the model, these properties most likely change with height in the ‘real world’. This is probably more important than merely lifting the clouds, and therefore it is not surprising that the modelled effect of raising the clouds is not clearly observed for the highest clouds on Fig. 9. * * * * 223 and the solar zenith angle. Some effects seem, however, to be due to other reasons. For Bergen, Heliosat provided too high estimates in the morning and too low estimates in the afternoon. The reason seems to be that Bergen is situated on the west coast, and hence clouds in the summer are more frequent to the east (over land) than to the west (over sea). For the co-scattering angle less than 5 we see that Heliosat does not account for the opposition effect; higher reflectivity due to less shadows leads to 5% too low global irradiances. However, this is the case for less than 2% of the hours, so the overall effect due to this is very small. Observed cloud amounts showed influence on the MBD, except for fully overcast, where Heliosat gave too high values for all three sites; 5% for Geneva and Lyon and 10% for Bergen. Two possible explanations are proposed; 1) less multiple reflection between ground and the atmosphere when there are no ‘holes’ in the cloud cover, and 2) clouds with full coverage may also be very thick. For all three sites, the height of the base of the lowest clouds showed a clear influence on the bias. This is also reproduced with a radiative transfer model, and a physical explanation is given. However, the effect is only seen for cloud bases below 600 metres, which includes only 12–18% of the hours, so the effect on the total bias is again small. 5. Conclusions Acknowledgements Global irradiances estimated with the Heliosat method using Meteosat-data from 1996=97 as input are compared to ground observed counterparts in Bergen, Geneva and Lyon. The biases are analysed with respect to variations of the sunground-satellite geometry and ground observed cloud amounts and height of cloud base. This work is a part of the project Heliosat-3 funded by the European Commission (NNK5-CT-200-00322). I thank project colleagues for valuable advice. Solar radiation and cloud data from Geneva was provided by Pierre Ineichen at CUEPE, University of Geneva and similar data from Lyon were provided by Dominique Dumortier at ENTPE, Lyon. Cloud observations from Bergen was provided by Magnar Reistad at the Norwegian Meteorological Institute. * Variations of MBD with solar zenith angle are small for clear cases. For cloudy cases Heliosat gives too low estimates for high sun and too high estimates for low sun. A possible explanation, which is supported by radiative transfer calculations, is that global irradiance varies differently with respect to solar zenith angle for clear and cloudy cases. Some variations of MBD with respect to the solar azimuth- and co-scattering angle are shown to be due to the relationship between these angles References Cano D, Monget JM, Albuisson M, Guillard H, Regas N, Wald L (1986) A method for the determination of the global solar radiation from meteorological satellite data. Solar Energy 37: 31–39 Dumortier D (1995) Mesure, Analyse et Modelisation du gisement lumineux Application a l’evalution des performances de l’eclairage naturel des b^atiments. PhD Thesis, Laboratoire Genie Civil et Habitat, ENTPE, Lyon, France Hammer A, Heinemann D, Westerhellweg A (1998) Derivation of daylight and solar irradiance data from satellite 224 K.-F. Dagestad: Mean bias deviation of the Heliosat algorithm observations. Proceedings of the 9th conference on Satellite Meteorology and Oceanography, Paris, May 1998, pp 747–750 Hammer A, Heinemann D, Hoyer C, Lorenz E (2003) Database of METEOSAT derived cloud indices for solar radiation. SODA deliverable D3–10 Iehle A, Bauer O, Wald L (1997) Final Report of ARMINES=ENSMP to the University of Oldenburg for the SATEL-LIGHT programme. Groupe teledetection et modelisation, ENSMP, Sophia Antipolis: France Ineichen P, Perez R (1999) Derivation of cloud index from geostationary satellites and application to the production of solar irradiance and daylight illuminance data. Theor Appl Climatol 64: 119–130 Leontieva E, Stamnes K, Olseth JA (1994) Cloud optical properties at Bergen (Norway) based on the analysis of long-term solar irradiance records. Theor Appl Climatol 50: 73–82 Olseth JA, Skartveit A (1993) Characteristics of hourly global irradiance modelled from cloud data. Solar Energy 51: 197–204 Olseth JA, Skartveit A (2001) Solar irradiance, sunshine duration and daylight illuminance derived from METEOSAT data for some European sites. Theor Appl Climatol 69: 239–252 Page J (1996) Algorithms for the Satellite programme. Technical Report for the second SATELLITE meeting in Bergen, Norway, June 1996 Radiation Yearbooks (1996–1997) Radiation observations in Bergen, Norway. University of Bergen, Norway Richiazzi P, Yang S, Gautier C, Sowle D (1998) SBDART: A research and teaching software tool for plane-parallel radiative transfer in the earth’s atmosphere. Bull Amer Soc 79: 2101–2114 Zelenka A, Perez R, Seals R, Renne D (1999) Effective accuracy of satellite-derived hourly irradiances. Theor Appl Climatol 62: 199–207 Author’s address: Knut-Frode Dagestad (e-mail: Knutfrode.Dagestad@gfi.uib.no), University of Bergen, Department of Geophysics, Allegaten 70, 5007 Bergen, Norway. Paper II Müller, R.W., Dagestad, K-F., Ineichen, P., Schroedter, M., Cros, S., Dumortier, D., Kuhlemann, R., Olseth, J. A., Piernavieja, C., Reise, C., Wald, L. and Heinemann, D. (2004) Rethinking satellite based solar irradiance modelling - The SOLIS clear sky module Remote Sensing of the Environment, 91 160-174 Remote Sensing of Environment 91 (2004) 160 – 174 www.elsevier.com/locate/rse Rethinking satellite-based solar irradiance modelling The SOLIS clear-sky module R.W. Mueller a,*, K.F. Dagestad b, P. Ineichen c, M. Schroedter-Homscheidt d, S. Cros e, D. Dumortier f, R. Kuhlemann g, J.A. Olseth b, G. Piernavieja h, C. Reise i, L. Wald e, D. Heinemann g a University of Oldenburg, now at Deutscher Wetterdienst, Frankfurter Street 135, 63067 Offenbach, Germany b University of Bergen, Bergen, Norway c University of Geneva, Geneva, Switzerland d German Aerospace Center—German Remote Sensing Data Center (DLR-DFD), Germany e Ecole des Mines de Paris, France f Ecole Nationale des Travaux Publics de l’Etat, France g University of Oldenburg, Oldenburg, Germany h Instituto Tecnologico de Canarias, Spain i Fraunhofer Institute for Solar Energy Systems, Germany Received 23 July 2003; received in revised form 26 February 2004; accepted 28 February 2004 Abstract Accurate solar irradiance data are not only of particular importance for the assessment of the radiative forcing of the climate system, but also absolutely necessary for efficient planning and operation of solar energy systems. Within the European project Heliosat-3, a new type of solar irradiance scheme is developed. This new type will be based on radiative transfer models (RTM) using atmospheric parameter information retrieved from the Meteosat Second Generation (MSG) satellite (clouds, ozone, water vapour) and the ERS-2/ENVISAT satellites (aerosols, ozone). This paper focuses on the description of the clear-sky module of the new scheme, especially on the integrated use of a radiative transfer model. The linkage of the clear-sky module with the cloud module is also briefly described in order to point out the benefits of the integrated RTM use for the all-sky situations. The integrated use of an RTM within the new Solar Irradiance Scheme SOLIS is applied by introducing a new fitting function called the modified Lambert – Beer (MLB) relation. Consequently, the modified Lambert – Beer relation and its role for an integrated RTM use are discussed. Comparisons of the calculated clear-sky irradiances with ground-based measurements and the current clear-sky module demonstrate the advantages and benefits of SOLIS. Since SOLIS can provide spectrally resolved irradiance data, it can be used for different applications. Beside improved information for the planning of solar energy systems, the calculation of photosynthetic active radiation, UV index, and illuminance is possible. D 2004 Elsevier Inc. All rights reserved. Keywords: Solar irradiance modelling; Remote sensing 1. Introduction Satellite-based remote sensing is a central issue in monitoring and forecasting the state of the Earth’s atmosphere. Geostationary satellites such as METEOSAT or GOES provide cloud information in a high spatial and temporal resolution. These satellites are, therefore, not only * Corresponding author. E-mail address: richard.mueller@dwd.de (R.W. Mueller). 0034-4257/$ - see front matter D 2004 Elsevier Inc. All rights reserved. doi:10.1016/j.rse.2004.02.009 useful for weather forecasting, but also for the estimation of solar irradiance, since the knowledge of the radiance reflected by clouds is the basis for the calculation of the transmitted irradiance. Additionally, a detailed knowledge about atmospheric parameters involved in scattering and absorption of sunlight is a further necessity. An accurate estimation of the downward solar irradiance is not only of particular importance for assessing the radiative forcing of the climate system, but also absolutely necessary for an efficient planning and operation of solar energy systems and the estimation of the energy load. Solar resource assessment R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 161 from geostationary satellites constitutes a powerful alternative to a meteorological ground network for both climatological and operational data (Perez et al., 1998). Solar irradiance schemes provide accurate solar irradiance data with a high spatial and temporal resolution using weather satellites such as METEOSAT and Meteosat Second Generation (MSG). Currently, most of the operational calculation schemes for solar irradiance are semiempirical and based on statistical methods. They use cloud information from the current METEOSAT or GOES satellite and climatologies of atmospheric parameters, e.g., turbidity (characterising the combined effect of aerosols and water vapour; see Perez et al., 2001 and references therein). The Heliosat method (Cano et al., 1986; Beyer et al., 2003) is certainly one of the best known. It converts METEOSAT satellite data into irradiance with a better accuracy than interpolated ground measurements could provide (Zelenka et al., 1999; Perez et al., 1998). It is applied routinely in real time at the University of Oldenburg since 1995. It has permitted the establishment of the server Satel-Light, which delivers valuable information on daylight in buildings to architects and other stakeholders (Fontoynont et al., 1997). It has also been used within the SoDa project1 (Wald et al., 2002) for the calculation of the solar irradiance. Furthermore there exists derivates of Heliosat, e.g., Heliosat-2 (Lefèvre et al., 2002), which is optimised as an operational processing chain for climatological data. With the launch of the Meteosat Second Generation (MSG) satellite, the possibilities for monitoring the Earth’s atmosphere have improved enormously. The MSG satellite will not only provide higher spatial (1 km) and temporal (15 min) resolution, but also offers with its 11 channels from 0.6 to 13 Am, the potential for the retrieval of atmospheric parameters such as additional cloud parameters, ozone, water vapour column, and with restrictions aerosols. These capabilities plus the synergy with other sensors, such as those aboard ERS-2 and ENVISAT (GOME/ATSR-2 and SCIAMACHY/AATSR), permit us to attain a refinement in the solar irradiance modelling. These refinements necessitate a rethinking of satellite-based solar irradiance modelling and going ahead with a drastic revision of the current Heliosat processing scheme. The current Heliosat scheme cannot exploit enhanced information about the atmosphere provided by improved satellite capabilities. Thus, it was necessary to develop a new scheme, which will be able to exhaust the enhanced capabilities of MSG (SEVIRI) and ENVISAT (SCIAMACHY). The accuracy of the calculated irradiance is expected to increase significantly with a scheme that can exhaust the capabilities of the new satellites. The new calculation scheme has to be fast, accurate, and should provide—in contrast to Heliosat and Heliosat-2—spectrally resolved solar irradiance data. As a consequence of the things mentioned above, the new scheme is based on the integrated use of a radiative transfer model (RTM), whereas the information of the atmospheric parameters retrieved from the MSG satellite (clouds, ozone, water vapour) and from the GOME/ATSR-2 instruments aboard the ERS-2 satellites (aerosols, ozone) will be used as input to the RTM-based scheme.2 The direct integration of an RTM into the calculation schemes— instead of using precalculated look-up tables—is only possible if the necessary computing time can be kept small. For this purpose, a functional treatment of the diurnal solar irradiance variation is applied, allowing an appropriate operational use of an RTM within the calculation scheme. This paper focuses on the description of the new clear-sky module, especially on the integrated use of the radiative transfer model (Section 2). The linkage of the clear-sky module with the cloud module is briefly described in order to point out the benefits of the integrated RTM use for allsky situations as well. 1 Integration and Exploitation of Networked Solar Radiation Databases for Environment Monitoring Project. 2 In the near future, the information from GOME/ATSR-2 will be replaced by SCIAMACHY/AATSR on ENVISAT. 2. SOLIS—the new scheme 2.1. Overview The integrated usage of the RTM within the scheme is related to the clear-sky scheme using the well-established n –k relation of the Heliosat method (Cano et al., 1986; Beyer et al., 1996) or the Cloud Optical Depth (COD) option to consider cloud effects. It is important to note that the integrated use of the RTM within the clear-sky module is linked with an enormous improvement for all-sky situations as well. It is not a restriction of the model. This issue will be discussed in more detail in Section 8.1. On the other hand, the benefits and needs of the described clear-sky module can only be understood if it is seen in the context of its main purpose—the operational satellite-based solar irradiance modelling with a large geographical coverage. Keeping this in mind, it is also necessary to describe briefly the treatment of the clouds and the basics of the linkage between the clearsky module—described in detail in this paper— and the cloud modules, which are partly still under development. The cloud modules will be discussed in more detail in a forthcoming paper after reliable MSG data will be available. 2.1.1. Using n – k relation The Heliosat method was originally proposed by Cano et al. (1986) and later modified by Beyer et al. (1996) and Hammer (2000). The basic idea of the Heliosat method is a two-step approach. In the first step, a relative normalised cloud reflectivity—the cloud index—is derived from METEOSAT images. The derived cloud index is correlated 162 R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 to the clear-sky index k, which relates the actual ground irradiance G to the irradiance of the cloud-free case Gclear sky. Consequently, in addition to the cloud index derived from the satellite signal, a clear-sky model, providing Gclear sky, is necessary for the estimation of the actual ground irradiance. The n – k relation is powerful, validated, and leads to small root mean square deviation (RMSD) between measured and calculated solar irradiance for almost homogenous cloud situations (RRMSD of 13 – 15% for hourly values; Hammer, 2000). With MSG data, it can be expected that the treatment of clouds using the current n– k will be improved only due to the higher spatial and temporal resolution. applied, making an appropriate explicit operational use of an RTM within the calculation schemes possible. Starting point of the integrated use is the assumption that daily values of the atmospheric clear-sky parameters in a spatial resolution of 100 100 or 50 50 km are sufficient. This assumption is reasonable for solar energy applications in consideration of accuracy and operational practicality, because of the following. 2.2. Basic considerations Daily values of water vapour and aerosols are linked with a great improvement compared to the current implicit use of a monthly turbidity climatology or aerosol and watervapor climatologies. Current restrictions in the art of retrieval limit the available input with respect to the temporal and spatial resolution of the atmospheric clear-sky parameters. For example, the retrieval of aerosols from satellites is handicapped by the small aerosol reflectance and the perturbation of the weak signal by clouds and surface reflection. In addition, the retrieval of water vapor is not possible for cloudy pixels. For these reasons, retrieval of daily values in 50 50-km resolution with a ‘‘global’’ coverage in an appropriate accuracy would be a great improvement. The effect of ozone is small compared to that of aerosols and water vapour; therefore, daily ozone values are sufficient. The temporal daily fluctuations of solar irradiance are generally dominated by cloud fluctuations. The cloud information is used in MSG pixel resolution (see Table 1), hence in a high temporal and spatial resolution. The usage of the modified Lambert – Beer function, described in Section 2.5, should enable the correction of derivations from the daily values of the clear-sky irradiance in an easy and fast manner (see Section 2.5). MSG will scan the atmosphere with a very high spatial resolution (see Table 1, e.g., approximately 2.5 million pixels have to be processed every 15 min for Europe). Thus, the computing time necessary to calculate the solar irradiance for each pixel has to be very small to make an operational usage of the solar irradiance scheme possible. Instead of using look-up tables, a new, more powerful and flexible method, the integrated use of RTM within the scheme based on a modified Lambert –Beer (MLB) relation, will be applied. The integration of an RTM into the calculation schemes, instead of using precalculated lookup tables, is only possible if the necessary computing time can be decreased enormously. For this purpose, a functional treatment of the diurnal solar irradiance variation was Using daily values of the atmospheric parameters (O3, H2O(g), aerosols) within a region of 100 100 km (50 50 km) and the modified Lambert – Beer function, only two RTM calculations are necessary to define the complete diurnal variation of the clear-sky irradiance for a given atmospheric state (see Fig. 1). The effect of clouds on the clear-sky irradiance is considered by using the n– k relation or the COD option (Section 2.1). By this way, the cloud effect is considered in MSG pixel resolution, whereby no additional explicit RTM runs are needed. Fig. 1 illustrates the new scheme and the integrated use of the RTM within the clear-sky scheme. The modified Lambert – Beer (MLB) function is discussed in detail in the next section. 2.1.2. Using COD-based code Within this option, the information of the cloud optical depth (COD) is used to consider the cloud effect. The COD will be retrieved operationally from MSG with software from the German Aerospace Center (DLR), based on the Apollo (Kriebel & Gesell, 1989; Saunders & Kriebel, 1988) or Nakajima (Nakajima & King, 1990) method. The RTM model SBDART (Ricchiazzi et al., 1998) has been used to find a parameterisation in order to relate the all-sky irradiance to the clear-sky irradiance. Within this parameterisation, the effective cloud-particle radii, derivable with the Nakajima and King (1990)-based scheme, can also be used. The derived parametrisation needs some fine-tuning and has to be tested with MSG data. It will be discussed in more detail in a forthcoming paper. Regardless of the treatment of clouds, the basis for the calculation of the all-sky radiation is the clear-sky module, which is described in detail in the next section. 2.3. The modified Lambert –Beer function Table 1 Improvements in METEOSAT resolution MSGDMETEOSAT The Lambert –Beer relation is given by Spatial resolution Temporal resolution Spectral channels 1/3 kmD2.5/5 km 15 minD30 min 12D3 I ¼ I0 expðsÞ ð1Þ R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 163 In a second step, a correction of s0, or the equivalent to this, of the parameter (s0)/(cos(hz)) is performed, leading to the so-called modified Lambert – Beer relation (MLB). s0 ð5Þ Iðhz Þ ¼ I0 exp cosðhz Þ cosa ðhz Þ The correction parameter a is calculated at a SZA of 60j. The modified Lambert – Beer relation (Eq. (5)) cannot only be applied to wavelength bands for direct irradiance, but also to wavelength bands of global and diffuse irradiance. However, in order to apply the modified Lambert – Beer relation to global and diffuse irradiance, the following things have to be considered. At low visibilities (high optical depth, high aerosol load), I0 in Eqs. (4) and (5) has to be enhanced for global and diffuse radiation. A general equation has been developed which is applied to I0 to get I0,enh. Idiffuse ð6Þ I0 I0;enh ¼ ð1 þ I0 Idirect Iglobal Fig. 1. Diagram of the spatial and temporal linkage between clear-sky and cloud information. where s is the optical depth and within the scope of atmospheric radiation, I is the direct radiation at ground with sun in zenith, I0 is the extraterrestrial irradiance. Consideration of path prolongation and projection to the Earth’s surface leads to Eq. (2), where hz is the solar zenith angle (SZA) and I(hz) is the irradiance at hz. s Iðhz Þ ¼ I0 exp cosðhz Þ cosðhz Þ ð2Þ This formula describes the behaviour of the direct monochromatic radiation in the atmosphere. Transformation of Eq. (2) leads to the optical depth s s ¼ ln Iðhz Þ I0 cosðhz Þ cosðhz Þ ð3Þ In order to apply the Lambert – Beer relation to wavelength bands of direct irradiance, the above Lambert – Beer relation (Eq. (2)) has to be applied in a modified manner, using a two-step approach. In a first step, the ‘vertical’ optical depth s0 is calculated at a SZA of zero degree (hz = 0), using Eq. (4), which is a special case of Eq. (3). Iðhz ¼ 0Þ s0 ¼ ln I0 ð4Þ In order to avoid switching between I0 and I0,enh, I0,enh is always used instead of I0 in Eqs. (4) and (5) for diffuse and global irradiance. Additionally, for diffuse irradiance, the multiplication with cos(hz) in Eq. (3) has to be skipped for diffuse irradiance, since consideration of the projection to the Earth’s surface is no longer feasible. It is important to notice that I(hz) in Eqs. (4) and (5) stands either for global, direct, or diffuse irradiance at the given SZA; the fitting parameter a has different values for direct, global, and diffuse irradiance; s0 is always calculated at a SZA of zero and the correction parameter a at a SZA of 60j, independent whether the MLB relation (Eq. (5)) is applied to global, direct, or diffuse irradiance. Using the Modified Lambert –Beer (MLB) relation, the calculated direct and global radiation can be reproduced very well (see Fig. 2). 2.3.1. General remarks The usage of the modified Lambert –Beer function is physically motivated, but it is actually a fitting function. This is especially obvious for the case of diffuse radiation. In principle, it is possible to fit the RTM calculations with any appropriate function, for example, a modified polynomial of third or higher degree (ecos3(x) + fcos2(x) + g). Hence, the big advantage of the modified Lambert –Beer function is not the feasibility to fit the RTM calculations, but that it is possible to yield a very good match between fitted and calculated values by using only two SZA calculations (e.g., better match than obtained with a polynomial of third degree). This is possible since the change of the irradiance 164 R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 with SZA is related to the Lamber –Beer law; hence, using the modified Lambert –Beer relation, ‘‘the degrees of freedom can be reduced.’’ Moreover, the parameter a can be calculated without the need for a numerical fit, respectively. The function was tested for many different atmospheric states, e.g., four different aerosol types, five different visibilities (5, 10, 23, 50, 100), different water-vapour amounts, different standard atmospheres, and surface models. No atmospheric clear-sky state is expected for which the MLB fit will not work. For our purpose, the sense of an appropriate fitting function is to save calculation time without losing ‘‘significant’’ accuracy. The question if a fitting function is usable for that purpose depends on the difference between the fitted values and the RTM calculated values. The differences in the broadband irradiance (306.8 – 3001.9 nm) are usually less than 8 W/m2 for high SZA and less than 5 W/m2 for SZA below 75j (8 and 4 W/m2 for direct irradiance, respectively). For the wavelength bands, the differences are typically less than 1 W/m2 below a SZA of 75j. 2.3.2. Discussion For monochromatic radiation, s is constant and consequently equals s0 for all SZA. In the case of wavelength bands, s is not constant, but changes smoothly with increas- Fig. 3. For monochromatic radiation, the Lambert – Beer relation is still a good approximation if the ‘vertical’ optical depth s0 is used. In order to yield a better match, a correction of formula (2) (MLB relation) is necessary. ing SZA. s0 is just the optical depth at hz = 0 and no longer equal to s for all SZA. The reason for that is the nonlinear nature of the exponential function; the monochromatic optical depths are, in contrast to the irradiance, not additive. I ¼ Iðk1 Þ þ Iðk2 Þ; but s psðk1 Þ þ sðk2 Þ ð7Þ That is the reason why a correction of the optical depth, or an equivalent to this, of the parameter (s)/(cos(hz)) is necessary. With respect to global irradiance, it has to be mentioned that the Lambert –Beer law (Eq. (2)) is still a good approximation for ‘monochromatic’ global radiation and moderate aerosol load, using s0 (see Fig. 3). For high aerosol load, the modified Lambert – Beer relation has to be used in order to get a good match with explicit RTM results. Moreover, for wavelength bands, the use of the modified Lambert – Beer relation is absolutely necessary. The Lambert – Beer relation describes the attenuation of incoming radiation. The incoming diffuse radiation at the top of the atmosphere is negligible. The source of diffuse radiation is the attenuation of the direct radiation due to scattering processes. Hence, the Lambert – Beer law is related to the amount of diffuse radiation, but does not describe the magnitude of the diffuse radiation. However, fitting with the modified Lambert – Beer relation works very well (see Fig. 2). 2.4. Radiative transfer model The radiative transfer model (RTM) used within the clear-sky module, LibRadtran,3 is a collection of C and Fortran functions and programs for calculation of solar and thermal radiation in the Earth’s atmosphere. It has been validated by comparison with other models (Koepke et al., 1998; Van Weele et al., 2000) and radiation measurements (Mayer et al., 1997). It is very flexible with respect to the Fig. 2. Comparison between RTM calculations and fit using the modified Lambert – Beer relation for different atmospheric states. 3 available at http://www.libradtran.org. R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 atmospheric input, e.g., different possibilities for the input of the aerosol information can be chosen by the user. LibRadtran offers the possibility of using the correlated-k approach of Kato et al. (1999). The correlated-k method is developed to compute the spectral transmittance (hence the spectral fluxes) based on grouping of gaseous absorption coefficients. The main idea is to benefit from the fact that the same value of the absorption coefficient k is encountered many times over a given spectral interval. Thus, the computing time can be decreased by eliminating the redundancy, grouping the values of k, and performing the transmittance calculation only once for a given value of k. Using the correlated-k option, the spectral resolved data can be calculated operationally in MSG pixel resolution, a new feature, so far not implemented in the Heliosat or Heliosat-2 method. Consequently, solar irradiance scheme (SOLIS) calculates the global, direct, and diffuse irradiance, not only for the broadband wavelength region (300 –300 nm), but for each of the Kato correlated-k (Kato et al., 1999) wavelength bands. The spectral output is provided for 27 bands between 306.8 and 3001.9 nm, which is sufficient for solar energy applications. Also, additional wavelength bands below 306.8 or above 3001.9 nm can be used. The MLB relation works very well for the spectrally resolved data (see Fig. 4 as an example). 2.5. Further benefits of the MLB function with respect to the use of water vapour, aerosol and ozone input information The modified Lambert –Beer relation is defined with the parameters ‘‘vertical’’ optical depth and the correction parameters ai. These parameters are calculated for a given atmospheric state (O3, H2O(g), AOD). A useful feature of the MLB is that different water vapour or ozone content affect the ‘‘vertical’’ optical depth s0 and not the correction parameter a (Fig. 5). Since s0 is calculated at a SZA of zero degree, the calculation and usage of look-up tables is 165 Fig. 5. Correction of H2O deviations. The values derived with the MLB relation are compared for different H2O(g) amounts. H2O is specified in ppm (parts per million). Recalculation of the vertical optical depth leads to a good agreement between MLB values and the explicit RTM runs for the same ai. straightforward, because calculations at zero degree SZA deal with vertical columns and not with slant columns. For aerosols, both quantities—s0 and ai—change for different values of the aerosol optical depth (AOD). In contrast to O3 and H2O(g), which are pure absorbers, aerosols are strong scattering particles, affecting not only s0 but also the correction parameter ai. If the changes in ai are neglected and only s0 is corrected, deviations occur increasing with increasing SZA. Consequently, another correction has to applied. In the case of aerosols, deviations from the assumed value can be approximately corrected by applying the following equations. To correct for increase of AOD from sA1 to sA2 sA1 Icor ¼ 2 IMLB I 0;sA2 I 0;sA2 cosðhz Þ I 0;sA1 ð8Þ To correct for decrease in AOD from sA2 to sA1 I 0;sA1 sA2 Icor ¼ 0:5 IMLB 0;s þ I 0;sA1 cosðhz Þ I A2 ð9Þ sA1 is the diurnal variation of the irradiance for the Here IMLB AOD A1 or A2, as given by the modified Lambert – Beer fit, I0,sA2 and I0,sA2 are the irradiances at a SZA of zero for AOD A1 and A2, respectively. Based on these equations (Eqs. (8) and (9)), the correction and use of look-up tables is Table 2 Ground stations used for the model – measurements comparison Fig. 4. Comparison between RTM calculations and fit using the modified Lambert – Beer relation. Example for a fit within a small wavelength band. Station Latitude (j) Altitude (m) Albany (NY) 42.7 100 Burns (OR) Eugene (OR) FSEC – Cocoa (FL) Geneva (CH) 43.6 44.1 28.3 46.2 1265 150 8 410 Climate Time base (min) humid continental semiarid temperate subtropical semicontinental 1 5 5 6 1 166 R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 Table 3 Aerosol load, water-vapor column, and Linke turbidity for the 13 considered days Station Day, year Tau 550 w (cm) TL Albany June 25, September 16, 2001 January 15, June 15, August 12, 2002 February 14, October 17, 2002 March 29, November 28, 1999 July 21, March 31, 1996, 1998 April 7, July 19, 2003, 2002 0.089, 0.048 0.027, 0.093, 0.053 0.073, 0.032 0.142, 0.081 0.093, 0.384 0.083, 0.087 3.0, 1.0 3.2, 2.5 2.0, 3.1, 2.5 2.7, 2.5 3.5, 3.0 3.0, 5.3 2.6, 3.0 Burns Eugene FSEC Geneva Fig. 6. Correction of Aerosol deviations, here for urban aerosols. The correction leads to an good agreement below SZA of 75j. The correction works also if marine aerosol with a AOD is corrected to urban aerosols with an AOD of 0.5. Geneva 0.4, 2.0, 1.0 1.1, 1.5 2.0, 2.0 1.5, 1.1 0.6, 1.7 3.1. Comparison of model results against ground measurements straightforward. Look-up tables, providing the irradiance for different AOD, has only to be calculated for a SZA of zero. All other information needed to apply the above correction equations are provided by the MLB fit, using daily values as input. By this way, the effect of deviation from the daily value can be corrected in an easy and fast manner. However, before the correction methods would be included in the operational SOLIS version, further validations and optimisation of the correction procedures have to be performed. 3. Intrinsic precision of the SOLIS irradiance In this section, the atmospheric data input is retrieved from the ground-based measurements used also for the model – solar irradiance data comparison. The main focus of this comparison is therefore to investigate the intrinsic precision of the direct-beam model. 3.1.1. Measurements The direct-beam and global irradiance produced with the SOLIS scheme is compared with ground measurements taken at five stations with different latitudes, altitudes, and climates as given in Table 2. The comparison is done against measurements for clear and stable meteorological conditions and for different water vapour and aerosol atmospheric loads (Fig. 6). From each of the five databases, 2– 4 clear days are extracted for winter and summer season. The stability of the atmospheric conditions is manually verified for each day: during the considered period of time, the water vapour content, the aerosol optical depth, and the Linke turbidity coefficient as defined by Ineichen and Perez (2002) are relatively stable as illustrated for February 14, 2002 in Eugene (OR) in Fig. 7.4 The days used in this comparison are listed in Table 3. Quality control has been done to eliminate specific measurements for which the direct-beam sensor is obstructed, but not the global sensor. 3.1.2. Retrieval of the atmospheric parameters For all the data, the Linke turbidity TL can be calculated from the normal direct-beam radiation. In order to ensure compatibility, even if the turbidity is relatively stable during the considered periods, the coefficient is evaluated at air mass 2. The water vapor column w is evaluated from ground measurements of the ambient temperature and relative humidity. With the knowledge of TL and w, and with the help of a model developed by Ineichen (2003), the aerosol optical depth can be retrieved. These three parameters are given in Table 3 for the considered data. An average value of 340 DU for the ozone content is taken for the comparison. It has been shown (Ineichen, Fig. 7. February 14, 2002, Eugene (OR). The stability of the atmospheric parameters is illustrated vs. solar time. 4 When a complete day is extracted, the morning/afternoon symmetry is respected. R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 167 IDMP station in Freiburg (47j59VN, 7j50VE) and with measurements of the meteorological station in Bergen, Norway (60.4jN, 5.3jE). Additionally, SOLIS calculations are compared with the clear-sky model used in the Heliosat method (Cano et al., 1986; Beyer et al., 2003). 4.1. Atmospheric data input The present study deals with the clear-sky case, relevant input parameter are ozone, water vapor, and aerosols. Standard climatology profitless are used in order to take the effect of Rayleigh scattering and the effect of other gas absorber into account. Fig. 8. Horizontal direct-beam irradiance evaluated by SOLIS vs. the correspondent ground measurements. 2003) that the influence of different ozone columns on broadband irradiance estimated by SOLIS is negligible. 3.1.3. Comparison The result of the comparison is given in Fig. 8. The graph illustrates the modelled horizontal direct-beam irradiance vs. the ground measurements. The mean bias difference between model and measurements (MBD) and the root mean square difference (RMSD) for the 4320 values are the following: horizontal direct-beam irradiance: MBD = 1 W/m2 or 0.2% and the RMSD = 11 W/m2, or 2.3% horizontal global irradiance: MBD = 0 W/m2 and the RMSD = 22 W/m2, or 4.0% The result for the global irradiance is remarkably good, taking into account that the same aerosol type was used for all the simulations. It is important to note that for the purpose of this comparison, the atmospheric clear-sky parameters are retrieved from the direct-beam measurements against which the model is compared. Within this scope, the result of the direct-beam comparison can be considered as the intrinsic precision of the SOLIS model, if accurate daily values of the atmospheric clear-sky parameter are used as input.5 4. Application of the model: comparison with measurements using autonomous atmospheric input The purpose of the comparison is to discuss the advantages and benefits of the SOLIS clear-sky module especially with respect to the use of aerosol and water-vapor information instead of turbidity. The benefits and limitations of the currently available atmospheric input date is discussed briefly as well. Therefore, calculation using the described SOLIS scheme are compared with measurements from the 5 Daily values are not daily means. 4.1.1. Ozone To derive actual distributions of total column ozone, backscatter measurements from the Global Ozone Monitoring Experiment (GOME) onboard the ERS-2 satellite are used (Burrows et al., 1998). The core element of the retrieval is a DOAS (Differential Optical Absorption Spectroscopy) fitting technique. Due to the scanning geometry, the level-2 total column ozone data are distributed heterogeneously in time and space. To gain synoptic distributions of total column ozone and to consider atmospheric variability the data assimilation technique Kalman-Filtering is used (Daley, 1991). It is applied in conjunction with a spectral statistical planetary wave approach (Bittner et al., 1997; Bittner & Erbertseder, 2000) For this study, GOME GDP level 2 data version 3.0 from ESA/DLR are used, this data are also available at http://wdc.dlr.de. The data assimilation approach delivers global ozone column maps for a certain point in time with a horizontal resolution of 0.36j. 4.1.2. Water vapour Total water-vapour column data (TWC) were prepared using the TOVS (TIROS Operational Vertical Sounder) instrument on the NOAA-14 satellite. TOVS raw data are analysed with the International TOVS Processing Package (ITPP; Jun et al., 1994; Jun, 1994), a physical retrieval scheme to derive atmospheric temperature and water vapour profitless for both cloudy and cloud-free situations. The average distance between retrievals is approximately 80 km, but the data are distributed irregularly in space (polar orbiting satellite). Therefore, a distance-weighting interpolation scheme is applied which delivers twice daily an European TWC data set with a spatial resolution of 0.5j (Schroedter et al., 2003). This data product is available at the World Data Center for Remote Sensing of the Atmosphere http://wdc.dlr.de). The quality of TWC in comparison with the ECMWF (European Center of Middle Range Weather Forecasting) model has been monitored for the whole year 2000. The comparison delivers differences from 0.19 F 4.41 mm6 for December 2000 to 4.56 F 5.75 mm for August 6 Bias F standard deviation. 168 R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 2000. This fits the required data accuracy of less than 10 mm very well. In the future, it is planned to use water vapour column information derived from MSG itself. This will provide an improved spatial resolution and a better coverage for both Europe and Africa. It can be expected that the accuracy of the retrieved H2O(g) amounts will be in the same range. 4.1.3. Aerosols The new surface irradiance scheme allows the use of further aerosol data sets with aerosol optical depth (AOD) and aerosol type as parameters. It has recently been shown that aerosol parameters can be retrieved over land from the MISR and MODIS instruments onboard the EOS-TERRA1 (launched December 1999) and EOS-AQUA (launched May 2002) satellites (Tanre et al., 2001; Kahn et al., 1997) and from a synergetic retrieval (Holzer-Popp et al., in press a,b) of SCIAMACHY and AATSR onboard the ENVISAT satellite (launched March 2002). Therefore, the proposed new scheme holds the potential to use this upcoming operational satellite data sets in order to include up-to-date aerosol information. In order to get the information about the aerosols for the comparison presented in this paper, Linke turbidity values based on Kasten (1996) together with the GADS/OPAC aerosol climatology are used. The GADS/OPAC climatology (Hess et al., 1998; Koepke et al., 1997) provides information of the AOD and the aerosol type, whereby the AOD is dependent on the relative humidity. However, the spatial and temporal resolution of the GADS/OPAC climatology is coarse, as only summer and winter season and a 5j spatial resolution is available. The Linke turbidity climatology provides Linke turbidities in a monthly and 5 min of arc angle resolution worldwide (available at http:// www.helioclim.net/-linke/index.html). It has to be noted that this small spatial resolution is just possible, using improved interpolation routines such as data fusion. The underlying measurement data have a much higher spatial resolution. averaging window. The input values for ozone and water vapor were 275 DU and 15 mm, respectively. Note that ozone has no big effect on the broadband irradiance, but does on the UV. The turbidity map provides a turbidity of 4 for the respective months. That corresponds to a visibility of 34 km and an aerosol optical depth (AOD) of 0.23, respectively. The conversion of turbidity to visibility has been performed with the radiative transfer model MODTRAN (Abreu & Anderson, 1996). GADS/OPAC provides an AOD of 0.18 –0.25 for relative humidities between 50% and 80% and urban as an aerosol type. The range of the AOD is in consistency with the visibility derived from the Linke turbidity climatology. The average relative humidity for the clear-sky days was approximately 50%, leading to an AOD of 0.18. In Figs. 9 and 10, the comparison between SOLIS calculated and measured direct and global irradiance is diagrammed. It has to be noted that whether urban or rural aerosols are used, no significant differences in the calculated direct solar irradiance occur. Hence, just the results for the urban aerosols are diagrammed. In the case of global irradiance, the chosen aerosol type has a significant effect on the global irradiance. In both figures, the results of the Heliosat clear-sky model, described in Beyer et al. (2003), are also diagrammed. Using the aerosol information provided by the OPAC/ GADS climatology (AOD of 0.18, urban aerosol type) as input, the calculated global and direct irradiance matches the measurements very well, as shown in Figs. 9 and 10. The relative root mean square error is 1.9% for global and 4.2% for direct irradiance with a relative bias of 0.6%and 0.5%, respectively. In addition, the SZA dependency is reproduced very well by the SOLIS model. In contrast to the results of the SOLIS calculations, the Heliosat model results in a good match for the global irradiance, but with a significant underestimation of the direct 4.2. Comparison of measurements and model, Freiburg, August 2000 Cloud-free situations were selected according to the cloud index derived with the Heliosat method from METEOSAT images. A situation was assumed to be cloud-free if the cloud index n of the respective pixel was within the interval from 0.03 to 0.03 and the spatial variation of the cloud index was less than 0.02. It is likely that some situations with partial cloud cover are still included, which especially affects the direct irradiance, leading to an increasing statistical uncertainty. The ground measurements have originally a temporal resolution of 10 s. They are averaged to 30-min means in accordance with the temporal resolution of the satellite. The point in time when the pixel above the measurement station is scanned from the satellite lies in the middle of the 30-min Fig. 9. Comparison between SOLIS and measurements using the GADS/ OPAC information for the aerosols. The calculated Heliosat clear-sky irradiance is also diagrammed. The differences between the models are mainly due to the different atmospheric input information. R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 169 Fig. 10. Comparison between SOLIS and measurements using the GADS/OPAC information for the aerosols. The calculated Heliosat clear-sky irradiance is also diagrammed. irradiance for the given turbidity of 4. Since the turbidity defines the attenuation of the direct irradiance, this indicates that the chosen turbidity is too low. However, decreasing the turbidity to values around 3 leads to a better match between the measurements and the Heliosat modelled direct irradiance, on the one hand, but it leads to an overestimation of the global irradiance on the other. The reason is the redundant information of the turbidity in comparison with a separated treatment of aerosol type, aerosol optical depth, and water vapour. The effect of the aerosol type on the global irradiance cannot be considered by using turbidity. Consequently, a consistent match between measurements and calculated direct and global irradiance is only possible using information about the aerosol optical depth, the aerosol type and the water content ‘‘separately.’’ Using the Heliosat clear-sky model or any other model that is just based on turbidity, the effect of different aerosol types on the global irradiance cannot be considered, because the information about the atmospheric state is redundant. This effect is even significant for the measurement site, but is higher for sites with higher aerosol load, or for sites characterised by special types of aerosols events, like desert storms or biomass burning. That is a drawback of Heliosat-1 and -2, but demonstrates the advantages of the SOLIS model. Moreover, reliable information of the spectral distribution of the irradiance cannot be derived by using only turbidity, without any additional information about the atmospheric state. Additionally, changes in stratospheric aerosols, e.g., an increase of the load after a volcanic eruption, cannot be treated with the current Heliosat method without a refitting of the empirical equation. Using SOLIS, just the enhanced aerosol load has to be changed in the input file and the effect is considered. In a comparison with measurements at Mannheim (Germany), it was possible to verify that urban aerosols with a AOD of 0.18 is a reliable input for Freiburg. The station Mannheim is nearby the station Freiburg and is characterised by a similar micro-climate—cities within the Rhine valley climate. The bias between SOLIS results and measurements was below 1x . 4.2.1. Spectral resolved irradiance data Using the same atmospheric input (urban aerosol, AOD = 0.18), the measured and calculated illuminance has been compared for August 2000, Freiburg. The illuminance is a measurement of a quantity of light as perceived by the human eye. In order to calculate the illuminance, the spectrally resolved irradiance output of SOLIS is weighted with the light sensitivity of the human eye. The derived value is then multiplied with 0.683 in order to convert W/m2 to klux. The measurements and the calculation matches very well, demonstrating that the spectral output of the model is reliable (see Fig. 11). In addition, the model results for rural aerosols are also diagrammed. Fig. 11. Measured and modelled illuminance, clear-sky situations, Freiburg, August 2000. 170 R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 Table 4 The clear days (all in May) used in this comparison Station Day Year Water column (mm) Ozone (DU) Bergen (60.4N, 5.3E) 7, 19 6, 8 12, 13, 14 1999 2000 2000 7.5, 10.6 8.1, 11.6 13.6, 11.0, 8.1 368, 316 347, 342 309, 306, 319 4.2.2. General remarks It has to be mentioned that a good match between measured and calculated irradiances cannot be expected for every month within one season. Similar comparisons for September 2000 and 1999 lead to an acceptable but inferior match between measurements and SOLIS calculated irradiance. This is most probably due to the fact that a seasonal aerosol climatology is to coarse, see also Section 4.3. Changes in the aerosol content within a season due to transport processes are not considered, but they affect the measurements. Using climatological atmospheric data instead of daily values as input to SOLIS increases the RMSE significantly. This effect has been investigated within a study for Geneva (P. Ineichen, personal communication). 4.3. Comparison in Scandinavia SOLIS calculations are compared against ground-measured hourly global and direct irradiance of the Nordic site: Bergen, Norway (60.4jN 5.3jE). Cloud-free days were selected according to the following criteria: (1) The ground-observed cloud cover should not exceed 1 octa during the day; (2) the curve of direct-beam normal radiation should be smooth and symmetric around noon solar time. To allow the comparison with the clear-sky model of Heliosat, which uses monthly Linke turbidity coefficients as input, days were selected from the same month (here May). The selected days are shown in Table 4. The value of the Linke turbidity for May is 3.6 for Bergen. Daily values of total water column (TWC) and ozone were used as input to SOLIS, see 4.1. The GADS/OPAC aerosol climatology suggests ‘maritime tropical’ and ‘summer maritime tropical’ as aerosol types for Bergen. In SOLIS, ‘maritime’ aerosols were used, in addition to ‘rural’ and ‘urban’ ones. For Bergen, the aerosol optical depth at 0.55 Am is given as 0.06 and 0.15 for a relative humidity of 10% and 90%, respectively. The actual relative humidity varies between 10% and 90% for the clear days, with the lowest humidities around and after noon. 4.3.1. Results 4.3.1.1. Direct irradiance. Fig. 12 shows the observed direct irradiance for the clear days in Bergen plotted against the SOLIS and Heliosat modelled values. Most observations fit well the SOLIS data for low and high aerosol optical depth (0.06 and 0.15). However, for 1 day in Bergen, the measured values are clearly lower and fit well with a higher aerosol optical depth of 0.28. This value is not unreasonable; Olseth and Skartveit (1989) found the range of aerosol optical depth in Bergen to be 0.07 to 0.28 with a mean value of 0.13. The effect of water vapour is seen to be smaller than the effect of aerosols. The effect of ozone is negligible (for broadband irradiance as here). However, for spectral output, also accurate values of water vapour and ozone are of great importance. It is clear that the Heliosat model with monthly input cannot be expected to be very accurate on day to day variations, but may give good mean values. It appears, however, that SOLIS is capable of reproducing hourly radiation, given that the daily input (in particular aerosols) is accurate. 4.3.1.2. Global irradiance. Observed global irradiances for Bergen are plotted against SOLIS and Heliosat modelled values in Fig. 13 (top). Here, the SOLIS values are too high, while the Heliosat clear-sky model is closer to the observed values. However, if the aerosol type is changed from ‘maritime’ to ‘urban’, SOLIS match the observed values much better (Fig. 13, bottom). For the calculated direct irradiance, the aerosol type was of little or no importance, but for global irradiance, the aerosol type (single scattering albedo) largely Fig. 12. Observed direct irradiances vs. modelled values by SOLIS and Heliosat. The aerosol optical depths are shown in the legend. Aerosol type is ‘‘maritime.’’ R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 171 Fig. 13. Same as Fig. 12, but for global irradiance. Aerosol type is ‘‘maritime’’ on upper figure, and ‘‘urban’’ on lower figure. affects the diffuse part. Since Bergen is a city with 235,000 citizens, urban aerosols are not unreasonable. The GADS/ OPAC climatology has quite coarse resolution (5j), so it cannot account for microclimatic conditions represented by cities, especially if they are surrounded by a rural region. 5. Discussion 5.1. Treatment of clouds The integrated use of the RTM is performed within the clear-sky module. For the treatment of clouds the n – k relation or the COD option is used. With respect to the COD option, an RTM is used to find the parameterisation, but not directly integrated. The question arises why an integrated use of an RTM within the cloud module is not needed on the one hand or not possible on the other. The n – k or COD option works well for almost homogenous cloud situations. Consequently, the difficulties or limitations of both options arises from heterogenous cloud effects. With respect to 3-D cloud effects, an operational usage of an RTM for the treatment of heterogenous clouds (whether directly or using precalculated look-up tables) is not feasible today. The limitations of 3-D cloud modelling do not enable realistic RTM calculations of 3-D cloud problems in an operational manner. Just case studies are feasible. With respect to the operational use of an RTM, the problem is the nonavailability of realistic specification of heterogenous clouds from measurements. MSG will not provide sufficient information about 3-D cloud characteristics. No other satellite or measurement setup provides this information for the needed temporal resolution and spatial coverage nowadays. Besides, an explicit or integrated use of RTM is not practicable since the needed calculation time of 3-D RTM models is too large for an operational adaptation. 5.2. Spectral resolved irradiance for cloudy situations Within previous studies, it has been shown that the cloud index is independent on the wavelength in the range of the MSG visible channels. However, the shape of the spectral clear-sky irradiance changes due to the scattering effects in consequence of the cloud particles. In order to correct the change in the spectral shape, the RTM model LibRadtran has been used to calculate look-up tables. First comparisons between measured and simulated illuminance indicate that the spectrally resolved output is now serviceable for ‘‘allsky’’ situations as well. However, further validations should be done. Using the COD option for the treatment of clouds, the effect of clouds on the spectral distribution of clear-sky irradiance is already considered. Consequently, the integrated use of the RTM is the basis for spectral resolved solar irradiance data for all-sky situations as well. 6. Summary and conclusion Within this paper, the power and advantages of the new SOLIS model have been discussed. The main scope has been the SOLIS clear-sky module, but also the treatment of the clouds, even though this is still under development,7 has been briefly discussed in order to explain the expected 7 However, a SOLIS all-sky working version is available on request. 172 R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 benefits of the integrated RTM use for the all-sky situations as well. The integration of the RTM into the calculation schemes is associated with a high flexibility with respect to changes of the atmospheric state and the different user requirements on the solar irradiance data. SOLIS provides the possibility to use enhanced information of the atmospheric state and, hence, the potential to improve the accuracy of the calculated direct, global, and diffuse irradiance. Additionally, spectrally resolved data can be calculated operationally in MSG pixel resolution. The Modified Lambert – Beer relation enables the integrated use of RTM within the clear-sky scheme. The integrated use of RTM is linked with high flexibility relating to the input of the atmospheric state, changes in theory (e.g., new aerosol models), and the desirable output parameters. The integrated use of the RTM is linked with the following benefits: Spectral information is automatically provided using the correlated-k option included in the RTM LibRadtran package (http://www.libradtran.org/). Consistent calculations of global, direct, and diffuse radiation for clear-sky cases within one single scheme considering different aerosol types and not only turbidity. Hence, an improved estimation of the relation between global and direct radiation is possible, especially for clear-sky situations. The separated use of H2O and aerosol information is a requirement for accurate information of the spectral distribution of irradiance. Deviations of the atmospheric state from the average (O3, H2O(g), aerosols) can be easily corrected using the results of the modified Lambert –Beer fit. Clear and easy linkage with cloudy sky scheme, whereby the treatment of heterogenous cloud effects is not restricted. The usage of the modified Lambert –Beer law enables not only the direct integration of an RTM into the irradiance scheme, but also the potential for the calculation and use of easy handling look-up tables. The advantages of the modified Lambert – Beer relation, and hence, the integrated RTM use, can be adapted to other solar irradiance models as well and is therefore a new milestone in solar irradiance modelling. The SOLIS model has been validated in three steps. In Section 2.3, it has been demonstrated that the MLB function matches the RTM results very well. The differences in the broadband irradiance (306.8 – 3001.9 nm) are usually less than 8 W/m2 for high SZA and less than 5 W/m2 for SZA below 75j (8 and 4 W/m2 for direct irradiance, respectively). In Section 3, it has been demonstrated that SOLIS is able to reproduce very accurately (hourly) values of direct and global broadband irradiance if accurate daily atmospheric input parameters are used. The estimated RMSD is 2.3% for direct irradiance and 4% for global irradiance. In Section 4, SOLIS has been compared with measurements and with the Heliosat method, whereby H2O(g) and O3 retrieved from satellite data and the GADS/OPAC aerosol climatology (Hess et al., 1998; Koepke et al., 1997) have been used as input to SOLIS. It has been demonstrated that SOLIS is able to reproduce the measurements very well within the scope of the uncertainties introduced by the rough spatial and temporal resolution of the aerosol climatology. For broadband clear-sky irradiance aerosols are by far the most important parameter. Since a model depends on accurate atmospheric input data, there is an urgent need to improve the information about aerosols. Additionally, it has been shown that the aerosol type has a significant effect on the global irradiance. A consistent match between model and measurements for both global and direct irradiance was only possible by consideration of the aerosol type. In contrast to the Heliosat method (Beyer et al., 2003), which is dedicated to use turbidity information, the full information about the clear sky atmosphere, including aerosol type, can be used by SOLIS. The improvements in the art of retrieval during the last years are impressive and still going on. Together with new remote sensing instruments such as SCIAMACHY or SEVIRI (aboard ENVISAT, MSG), it can be expected that the information about the atmospheric state, especially with respect to clouds and aerosols, will be further improved. In order to benefit from the enhanced information about the atmospheric state, a model such as SOLIS is necessary. Keeping this in mind, the final conclusion can be drawn that the enhanced capabilities of the new MSG and ENVISAT satellites, together with the new type of solar irradiance scheme, SOLIS, will provide solar irradiance data with high accuracy, high spatial and temporal resolution, and large geographical coverage, all within the European Heliosat-3 project. Acknowledgements The Heliosat-3 project is funded by the EC (NNK5-CT200-00322).We thank Arve Kylling (NILU) and Bernhard Mayer (DLR) for providing the LibRadtran RTM package. The DWD is acknowledged for the data of the Mannheim station. Thanks to A. Drews for the grammatical corrections. References Abreu, L., & Anderson, G. (1996). The MODTRAN 2/3 report and LOWTRAN model. Technical report, Philips Laboratory, Hanscom. R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 Beyer, H., Costanzo, C., Heinemann, D. (1996). Modifications of the heliosat procedure for irradiance estimates from satellite images. Solar Energy, 56, 207 – 212. Bittner, M., & Erbertseder, T. (2000). The STREAMER project: An overview. Air pollution VIII. Southhampton, Boston: WIT Press. Bittner, M., Dech, S., Meisner, R., & Ruppert, T. (1997). Generating GOME level 3 data products at DFD. Technical report, DLR-DFD. Burrows, J., Weber, M., Buchwitz, M., Rozanov, V., Ladstätter-Weißenmayer, A., Richter, A., de Beek, R., Hoogen, R., Bramstedt, K., Eichmann, K.-U., Eisinger, M., & Perner, D. (1998). The Global Ozone Monitoring Experiment (GOME): Mission concept and first scientific results. Journal of Atmospheric Sciences, 56, 151 – 175. Cano, D., Monget, J., Albuisson, M., Guillard, H., Regas, N., & Wald, L. (1986). A method for the determination of the global solar radiation from meteorological satellite data. Solar Energy, 37, 31 – 39. Daley, R. (1991). Atmospheric data analysis. Cambridge, UK: Cambridge University Press. Fontoynont, M., Dumortier, D., Heinemann, D., Hammer, A., Olseth, J., Skartveit, A., Ineichen, P., Reise, C., Page, J., Roche, L., Beyer, H., & Wald, L. (1997). Satellight: A european programme dedicated to serving daylight data computed from meteosat images. Proceedings Lux Europa 1997. Vol. The 8th European Lighting Conference, Amsterdam. Available at: http://www.satellight.com/ indexgT.htm. Hammer, A. (2000). Anwendungsspezifische Solarstrahlungsinformationen aus Meteosat-Daten. PhD, School of Mathematics and Natural Sciences. University of Oldenburg. Hammer, A., Heinemman, D., Hoyer, C., Kuhlemann, R., Lorenz, E., Müller, R., & Beyer, H. G. (2003). Solar energy assessment using remote sensing technologies. Remote Sensing of the Environment, 86(3), 423 – 432. Hess, M., Koepke, P., & Schult, I. (1998). Optical properties of aerosols and clouds: The software package opac. Bulletin of the American Meteorogical Society, 79, 831 – 844. Holzer-Popp, T., Schroedter, M., & Gesell, G. (2002a). Retrieving aerosol optical depth and type in the boundary layer over land and ocean from simultaneous GOME spectrometer and ATSR-2 radiometer measurements: 1. Method description. Journal of Geophysical Research, 107(D21), 4578. Holzer-Popp, T., Schroedter, M., & Gesell, G. (2002b). Retrieving aerosol optical depth and type in the boundary layer over land and ocean from simultaneous GOME spectrometer and ATSR-2 radiometer measurements: 2. Case study application and validation. Journal of Geophysical Research, 107(D24), 4770. Ineichen, P. (2003). Broadband comparison of SOLIS clear sky scheme against other clear sky models and 5 data banks. Working paper Heliosat-3 project. Ineichen, P., & Perez, R. (2002). A new airmass independent formulation for the Linke turbidity coefficient. Solar Energy, 73, 151 – 157. Jun, L. (1994). Temperature and water vapour weighting functions from radiative transfer equation with surface emissivity and solar reflectivity. Advances in Atmospheric Sciences, 11(4), 421 – 426. Jun, L., Fengxian, Z., & Qingcun, Z. (1994). Simultaneous non-linear retrieval os atmospheric temperature and absorbing constituent profiles from satellite infrared sounder radiances. Advances in Atmospheric Sciences, 11(2), 128 – 138. Kahn, R., West, R., McDonald, D., Rheingans, B., & Mishchenko, M. (1997). Sensitivity of multiangle remote sensing observations to aerosol sphericity. Journal of Geophysical Research, 102, 16861 – 16870. Kasten, F. (1996). The Linke turbidity factor based on improved values of the integral Rayleigh optical thickness. Solar Energy, 56, 239 – 244. Kato, S., Ackerman, T., Mather, J., & Clothiaux, E. (1999). The k-distribution method and correlated-k approximation for a short-wave radiative transfer. Journal of Quantitative Spectroscopy and Radiation Transfer, 62. 173 Koepke, P., Bais, A., Balis, D., Buchwitz, M., de Backer, H., de Cabo, X., Eckert, P., Eriksen, P., Gillotay, D., Koskela, T., Lapeta, V., Litynska, Z., Lorente, J., Mayer, B., Renauld, A., Ruggaber, A., Schauberger, G., Seckmeyer, G., Seifert, P., Schmalwieser, A., Schwander, H., Vanicek, K., Weber, M. (1998). Comparison of models used for UV index calculations. Photochemistry and Photobiology, 67, 657 – 662. Koepke, P., Hess, M., Schult, I., & Shettle, E. (1997). Global aerosol data set. Technical report, MPI Meteorologie Hamburg Report. Kriebel, K. T., Saunders, R., & Gesell, G. (1989). Optical properties of clouds derived from fully cloudy AVHRR pixels. Beiträge zur Physik der Atmosphäre, 62, 165 – 171. Lefèvre, M., Albuisson, M., & Wald, L. (2002). Joint report on interpolation scheme ‘meteosat’ and database ‘climatology’ i (meteosat). Technical report, Report for the European Commission. Available at http:// www.soda-iscom/publications. Mayer, B., Seckmeyer, G., & Kylling, A. (1997). Systematic long-term comparison of spectral UV measurements and UVSPEC modeling results. Journal of Geophysical Research, 102, 8755 – 8767. Nakajima, T., King, M. (1990). Determination of the optical thickness and effective particle radius of clouds from reflected solar radiation measurements: Part I. Theory. Journal of the Atmospheric Sciences, 47(15). Olseth, J., & Skartveit, A. (1989). Observed and modelled hourly luminous efficacies under arbitrary cloudiness. Solar Energy, 42, 221 – 233. Perez, R., Aguiar, R., Collares-Pereira, M., Dumortier, D., EstradaCajigal, V., Gueymard, C., Ineichen, P., Littlefair, P., Lund, H., Michalsky, J., Olseth, J., Renne, D., Rymes, M., Skartveit, A., Vignola, F., Zelenka, A. (2001). Solar resource assessment: A review. Solar energy—the state of the art. No. ISBN 1 902916239 in ISES position papers. ( pp. 497 – 562). London: James and James Science Publishers. Perez, R., Renne, D., Seals, R., & Zelenka, A. (1998a). The strength of satellite-based solar resource assessment. Production of site/timespecific irradiances from satellite and ground data. Report 98-3. New York State Energy Research and Development Authority, Corporate Plaza West, 286 Washington Evenue Extension, Albany, NY 12203-6399. Perez, R., Seals, R., & Zelenka, A. (1998b). Production of site/time-specific hourly irradiances—Satellite remote sensing vs. network interpolation. Production of site/time-specific irradiances from satellite and ground data. Report 98-3. New York State Energy Research and Development Authority, Corporate Plaza West, 286 Washington Evenue Extension, Albany, NY 12203-6399. Ricchiazzi, P., Yang, S., Gautier, C., & Sowle, D. (1998). Sbdart: A research and teaching software tool for plane-parallel radiative transfer in the earth’s atmosphere. Bulletin of the American Meteorological Society, 79(10). Saunders, R., & Kriebel, K. (1988). An improved method for detecting clear sky and cloudy radiances from AVHRR data. International Journal of Remote Sensing, 9, 123 – 150. Schroedter, M., Olesen, F., & Fischer, H. (2003). Determination of land surface temperature distributions from single channel IR measurements: an effective spatial interpolation method for the use of TOVS, ECMWF and radiosonde profiles in the atmospheric correction scheme. International Journal of Remote Sensing, 24(6), 1189 – 1196. Tanre, D., Kaufman, Y., Holben, B., Chatenet, B., Karnieli, A., Lavenu, F., Blarel, L., Dubovik, O., Remer, L., & Smirnov, A. (2001). Climatology of dust aerosol size distribution and optical properties derived from remotely sensed data in the solar spectrum. Journal of Geophysical Research, 106(18). Van Weele, M., Martin, T., Blumthaler, M., Brogniez, C., den Outer, P., Engelsen, O., Lenoble, J., Pfister, G., Ruggaber, A., Walravens, B., Weihs, P., Dieter, H., Gardiner, B., Gillotay, D., Kylling, A., Mayer, B., Seckmeyer, G., & Wauben, G. (2000). From model intercomparisons towards benchmark UV spectra for six real atmospheric cases. Journal of Geophysical Research, 105, 4915 – 4925. 174 R.W. Mueller et al. / Remote Sensing of Environment 91 (2004) 160–174 Wald, L., Albuisson, M., Best, C., Delamare, C., Dumortier, D., Gaboardi, E., Hammer, A., Heinemann, D., Kift, R., Kunz, S., Lefèvre, M., Leroy, S., Martinoli, M., Ménard, L., Page, J., Prager, T., Ratto, C., Reise, C., Remund, J., Rimoczi-Paal, A., Van der Goot, E., Vanroy, F., & Webb, A. (2002). Soda: A project for the integration and exploitation of networked solar radiation databases. In W. Pillmann, & K. Tochtermann (Eds.), Environmental communication in the information society: No. Part 2 ( pp. 713 – 720). Vienna, Austria: International Society for Environmental Protection. Zelenka, A., Perez, R., Seals, R., & Renne, D. (1999). Effective accuracy of satellite derived hourly irradiances. Theoretical and Applied Climatology, 62, 199 – 207. Paper III Dagestad, K-F. and Olseth, J.A. (2005) An alternative algorithm for calculating the cloud index Manuscript An alternative algorithm for calculating the cloud index Knut-Frode Dagestad and Jan Asle Olseth Geophysical institute University of Bergen Allegaten 70 5009 Bergen Norway April 2005 Abstract The cloud index is an important component of the Heliosat algorithm, which estimates solar radiation components from Meteosat High Resolution Visible images. The cloud index quantifies the reflective properties of the atmosphere, and varies from 0 at clear conditions to 1 at overcast. The algorithm is semi-empirical in the way that it includes several constants that need to be tuned. Some of these were removed in the Heliosat-II algorithm (Rigollier et al., 2004) which introduced the Meteosat calibration constant to replace the "pseudo reflectivity" with a "real reflectivity". This approach is followed here, and two additional changes are made: 1) An analytical expression is introduced to correct for backscattered radiation from air molecules. 2) A correction is made for non-lambertian reflectivity, removing the time consuming need for determining the ground reflectivity for each month and each slot. The new cloud index is used to calculate global irradiances which are validated against hourly measurements from five European ground stations. The average root mean square deviation is 15.5% for a six month spring/summer period, of comparable accuracy as using the more time consuming traditional algorithm of Hammer (2000). 1 Introduction For more than 20 years global irradiance at ground level has been successfully estimated from images taken by meteorological satellites. Since the input to the models is extremely simple, a single digital count per pixel, the methods need also to be very simple. So instead of a physical approach of using forward calculations and the principle of conservation of energy, the methods rely on the simple idea of using relative values of the digital counts: when a pixel is relatively dark, cloud free conditions are assumed, and the output of the model is simply global irradiance calculated with a clear sky model when the pixel is relatively bright, overcast conditions are assumed, and the output of the model is e.g. 5% of the corresponding clear sky value for intermediate brightness a simple linear transformation is assumed On top of this scheme empirical corrections have been made with success, e.g. a subtraction of the scattered radiance from air molecules, which depends strongly on the geometry. Although such corrections improve the performance, they have some disadvantages: it is less obvious how to interpret physically the "relative reflectivity" and the cloud index 1 corrections based on tuning to data may have unpredictable effects outside the specific sites or time periods which they are tuned to will new corrections describe a real physical phenomenon or just a side effect of earlier corrections? Therefore, the number of corrections should be kept to a necessary minimum, and based on physical reasoning whenever possible. Here an analytical expression for the backscattered radiation from the atmosphere is derived and applied to the Heliosat-scheme. This removes the uncertainty related to the empirical expression from Hammer (2000) which could include components from aerosols and the sea surface, and which was also tuned to measurements for certain pixels for a certain time period. The analytical expression is also more straightforward to adapt to other sensors. Another necessary procedure in the Heliosat-scheme is to compensate for the lower bound of the relative reflectivity (ρground) which varies with time. The current approach in the Heliosat-scheme uses a histogram technique to determine the ground reflectivity for each slot (images acquired at the same UTC-time of day belong to the same "slot") and month, thus keeping the sun-ground-satellite geometry fairly constant. Some problems with this approach are: The number of data points available to find ρground is maximum 31, which sometimes gives numerical unstabilities This procedure is a very time-consuming part of the Heliosat scheme Besides, even for the same time of day, the geometry can change somewhat during a month. For Paris, as an example, the angle between the directions towards the sun and Meteosat for the 12 UTC slot is varying between 4 and 16 degrees during September. To overcome these problems, ρground is parameterised as a function of the angle between the directions to the sun and satellite as seen from ground ("co-scattering angle"). In addition to saving computing time, this permits the use of a much lager data sample to determine ρground , thus eliminating the problems for pixels and slots with few clear situations during a month. 2 An alternative algorithm for the cloud index 2.1 Calculation of the reflectivity from Meteosat counts A part of the signal that a "visible" satellite sensor receives when viewing earth is directly scattered from air molecules. This part depends strongly on the sun-ground-satellite geometry, and as the radiation at ground is independent of the satellite position, it should be corrected for. The traditional approach in the Heliosat algorithm is to subtract a quantity from an expression which is tuned to satellite counts from cloud free pixels over sea (Hammer, 2000). It was however shown by Dagestad (2001) that most of this signal is first order scattered radiance, and hence an analytical expression for this component could be used. Under the assumption of a plane-parallel atmosphere the following expression for radiance scattered towards a satellite is derived (see Appendix): 1 2 1 − 31cos cos r atm =I 0 [1−e cos cos ] 16 cos cos (1) where θ is the solar zenith angle, φ is the satellite zenith angle and ψ is the "co-scattering angle". I0 is the solar constant of 1367 W/m2. According to the Appendix an optical depth τ of 0.0426 is representative for the Meteosat-7 and 8 HRV channels, corresponding to an "equivalent wavelength" of 680 nanometres. Equation 1 is singular for θ or φ at 90 degrees, but should have sufficient accuracy up to at least 85º 2 for a spherical atmosphere. The advantage of this expression compared to the one from Hammer (2000) is that it is not fitted to certain angular configurations, and that it contains no signal from the surface or other atmospheric components. The reflectivity is then calculated by: = C−C off c f r atm − I cos I 0 cos (2) where: - C is the raw Meteosat HRV counts • - Coff is the constant instrument offset (51 for Meteosat-8) • - cf = 0.56 W W is the calibration constant and Iμ = 1403 2 is the m ⋅str⋅ m⋅counts m ⋅m band • 2 solar irradiance of the Meteosat-8 HRV channel (Govaerts et al., 2004) - ε is the correction for varying sun-earth distance The factor π is included to convert the reflected radiance to irradiance under the assumption of lambertian reflectance. This assumption is discussed in the next section. For the calculation of the cloud index ρ could be interpreted as the reflectivity of the ground and clouds, although this is not strictly physical correct. 2.2 Calculation of the cloud index from the reflectivity In the Satel-Light version of the Heliosat scheme (Fontoynont et al., 1998) the (pseudo) ground reflectivity is determined for each pixel and for each month and slot. It is however seen that the reflectivity depends strongly on the co-scattering angle ψ, and thus a parameterisation will be made to correct for this. The correction probably includes the effects of: non-lambertian reflection from the ground surface itself varying amounts of shadows due to nearby terrain and broken clouds scattering and absorption due to interaction with air molecules, aerosols and clouds Figure 1 shows the reflectivity according to Equation 2 plotted versus ψ for six sites in Europe and the Canary Islands. For each of these sites the 4-percentile value is calculated for ψ within each ten degree bin. A 3rd order polynomial is then fitted to these points, and plotted as broken curves on the figure. The mean of the polynomials is taken, and normalised to the value 1 for ψ = 0˚, to be used as a "shape function": g shape =1−0.59 0.11 20.05 3 (3) where ψ is given in radians. The ground reflectivity can then be estimated by: ground = g0 g shape (4) where ρg0 is the reflectivity of the pixel for ψ=0. This constant is determined by taking the 4percentile of a time series of reflectivities divided by the "shape function". To avoid noise ψ should be kept below 50˚. The advantage of this approach is that the ground reflectivity can be determined once and for all, saving a lot of computer power. Besides, the difficulty of determining the reflectivity for months/slots with few clear situations is also avoided. ρground from Equation 4 is plotted as solid lines on Figure1. Still the ground albedo can be determined more frequently to account for effects which are truly due to changes of the reflecting properties of the ground surface (e.g. snow cover and vegetative changes). 3 Figure 1: Reflectivities for the Meteosat-8 pixels of 6 European sites (dots) calculated with Equation 2 for all Meteosat8 images between 16 March and 31 August 2004, plotted as a function of the co-scattering angle ψ. Broken lines: third order polynomials of ψ fitted to 4 percentiles within each ten degree bin (0-10, 10-20, etc). Solid lines: ground albedo calculated by Equation 4 by use of the procedure described in section 2.2 The upper boundary of reflectivities (cloud reflectivity) is seen to vary much less with ψ (or solar zenith angle θ) and a constant value of 0.81 is chosen as ρcloud, taken as the 98 percentile of the counts. This is assumed to be the reflectivity of the "thickest clouds". The 98 percentile is chosen instead of the maximum value to avoid any outliers. The cloud index (n) is then finally calculated from: n= − ground cloud − ground 3 Description of data for validation 3.1 Ground measured global irradiances In this study the satellite derived irradiances are compared to hourly measurements of global irradiances at five European stations for the period 16 March to 31 August 2004 (Table 1). All measurements are done with Kipp & Zonen pyranometers, and the data are manually quality controlled by the respective data providers (see acknowledgements). 4 (5) Table 1: Stations with ground based hourly global irradiances. Station Elevation Latitude Longitude [m.a.s.l.] [ºN] [ºE] Instrument Barcelona 98 41.39 2.12 Kipp & Zonen CM 11 Bergen 45 60.40 5.32 Kipp & Zonen CM 11 Freiburg 275 48.02 7.84 Kipp & Zonen CM 11 Geneva 425 46.20 6.13 Kipp & Zonen CM 10 Lyon 170 45.78 4.93 Kipp & Zonen CM 6 3.2 Satellite derived global irradiances Global irradiances are calculated by the following empirical relationship (Rigollier et al., 1998) using the cloud indices of section 2.1 as input: 1.2 1−n k= 2 2.0667−3.6667 n1.6667 n 0.05 for n −0.2 for n∈[−0.2 , 0 .8] for n∈[ 0.8 , 1.1] for n1.1 (6) The clear sky index k is defined as the ratio between the actual global irradiance, G, and the clear sky global irradiance, Gclear, which can be modelled with a clear sky model. k≡ G G clear (7) Two different clear sky models are used in this study: The model used by the Satel-Light project (Fontoynont et al., 1998) which consists of one model for the direct irradiance (Page, 1996) and one model for the diffuse irradiance (Dumortier, 1995). The input used is height above sea level, solar elevation and monthly values of Linke turbidities from a database developed by Dumortier (1998). The SOLIS model (Mueller et al., 2004) which uses two simulations with the radiative transfer model libRadtran (www.libRadtran.org) per day to parameterize the diurnal variation of global irradiance (and other spectral and angular components). SOLIS uses climatological values of water vapour (NVAP, www.stcnet.com/projects/nvap.html) and aerosols (SYNAER, Holzer-Popp et al., 2002a, 2002b) as input. Operational retrieval of aerosols and water vapour for input to SOLIS is planned for the near future within the EUproject Heliosat-3. Two different cloud indices are also used: the cloud index calculated with the algorithm in section 2 the "old" cloud index described in Hammer (2000). Adaptation from Meteosat-7 to Meteosat-8 was performed by Annette Hammer and Rolf Kuhlemann at the University of Oldenburg (personal communication) including a change of ρcloud from 160 counts to 650 counts (Meteosat-8 gives 10-bit values, whereas Meteosat-7 gives 8-bit values) 5 4 Validation 4.1 Verification of the algorithm for calculation of ground reflectivity In section 2.2 a new method to calculate the ground reflectivity in Heliosat was developed. This method will here be validated against five stations which are independent from the stations used for development. Figure 2 shows reflectivities calculated with Equation 2 from the hourly means of the satellite counts for the five stations in table 1. It is seen that the ground albedo calculated with the algorithm of section 2.2 is nicely fitting the lower bound of reflectivities. The fit is actually better than for the development stations; the reason for this is that the averaging of four 15 minute values to create hourly values is removing much of the "noise" seen in the plot of 15-minute values on Figure 1. The shape function (Equation 3) describes well the non-lambertian variation of reflectivity for all stations, even though ρg0 varies between 0.165 for Bergen and 0.208 for Barcelona and Lyon. Figure 2: Calculated reflectivities for the hourly means of the satellite counts for the stations of Table 1 (points). Ground albedo from the algorithm in section 2.1 is plotted as solid lines. 4.2 Optimal pixel size In earlier versions of Heliosat, with data from Meteosat First Generation (MFG, Meteosat 1-7), the best accuracy was obtained by averaging cloud indices over 15 pixels, 5 pixels in the east -west direction and 3 pixels in the north-south direction (Fontoynont et al., 1998). To find the optimal pixel size for Meteosat-8, cloud indices are here calculated for the following configurations: single pixel, 3x3, 5x5, 7x7 and 1x3, 3x5, 5x7 and 7x9. The n times (n+2) configurations gives an approximately square area in Europe where pixels are longer in the north-south direction due to 6 oblique viewing angle. Figure 3 shows the Root Mean Square Deviation (RMSD) of hourly global irradiances for all five ground stations. Only results with the Satel-Light clear sky model are shown, but both the old and the new cloud indices are used. For both cloud indices and for all stations, averaging over 3x5 pixels gives the lowest RMSD. (The only exception is Lyon where 5x5 pixels gives slightly lower RMSD). This is the same results as for MFG, even though the MSG pixels have roughly nine times smaller area. In this study all irradiances are hereafter calculated with cloud indices averaged over 3x5 pixels. Figure 3: Root Mean Square Deviation (RMSD) of hourly global irradiances calculated with the Heliosat-method, using two different cloud indices. For the solid lines the cloud indices are averages over n x n pixels; for the broken lines the number of pixels in the east-west direction is n+2. 4.3 Validation of hourly global irradiances Table 2 shows RMSD and Mean Bias Deviation (MBD, model-observation) for all stations and for both cloud indices and both clear sky models (chapter 3). Only hours for which the solar elevation is always above 5 degrees are compared. For the average over all stations the new cloud index with the Satel-Light clear sky model gives the lowest RMSD (15.5%) and also the smallest MBD (-0.9%). However, when the SOLIS clear sky model is used the old cloud index performs better, and the global irradiances calculated with the new index are then on average 3% too low. 7 Table 2: Root Mean Square Deviation (RMSD) and Mean Bias Deviation (MBD, model-observation) for two different clear sky models (SOLIS, Satel-Light) and two different cloud indices ("New" and "Old", see chapter 3). The values are given in percent of mean observed global irradiance. Bold numbers show the lowest RMSD for the given station. New cloud index SOLIS Observed Station [W/m2] Old cloud index Satel-Light SOLIS Satel-Light Number of RMSD MBD RMSD MBD RMSD MBD RMSD MBD hours Barcelona 531 1475 14.7 -6.4 12.9 1.1 13.7 -3.9 13.6 3.8 Bergen 280 2073 23.4 0.2 23.5 0.4 23.1 0.0 23.3 0.2 Freiburg 389 1753 17.1 -6.0 16.3 -2.2 16.5 -3.3 16.5 0.6 Geneva 459 1656 14.0 -3.1 13.9 -2.3 13.8 -0.6 13.9 0.2 Lyon 432 2003 13.1 0.4 13.3 -1.3 13.6 2.5 13.4 0.8 All stations 410 8960 16.1 -3.0 15.5 -0.9 15.7 -1.0 15.7 1.2 Table 2 gives no unique answer to which is the best cloud index and which is the best clear sky model. According to Table 3 the Satel-Light clear sky model generally gives higher clear sky irradiances than the SOLIS model, and Figure 4 shows that the new cloud index is generally higher than the old one. Hence a too high cloud index can compensate for a too high clear sky value and vice versa. The frequency distributions of Figure 4 show that the new cloud index has a clear peak close to zero, while the old one has more frequently negative values. Negative cloud indices give clear sky indices higher than 1, which could compensate for a clear sky model giving too low values. For the clear sky models it should be stressed that the input to the models is probably more important than the models themselves, so an interesting prospect for the near future is the inclusion of daily retrieved atmospheric parameters into the SOLIS model within the Heliosat-3 project. The new cloud index has however the advantage of being fast and easy to calculate routinely, and as it is based on physical quantities it should be easier to do possible physical and/or empirical corrections in the future. Table 3: Mean clear sky global irradiances of all hours for all stations and both clear sky models. See chapter 3 for description of the models. SOLIS Satel-Light Barcelona 613.1 662.2 Bergen 484.2 484.6 Freiburg 576.9 600.5 Geneva 605.9 610.8 Lyon 606.2 595.5 All stations 573.3 584.6 8 Figure 4: Frequency distributions of the two different cloud indices for the five stations of Table 1. 5 Summary and discussion Two modifications are made to the traditional algorithm for calculation of the cloud index: An analytical expression for the scattered radiance from air molecules is introduced, replacing an empirical expression. The variation of the parameter ρground is parameterised as a function of the angle between the directions towards the sun and the satellite. This permits to calculate ρground once and for all, replacing a time consuming histogram technique which has to be applied separately to each slot and month. These two corrections are applied in the Heliosat scheme to calculate global irradiances and the results are compared to hourly measurements from five ground stations in Europe. The RMSD and the MBD are very similar compared to the results using the traditional cloud index from Hammer (2000). One cloud index gives however better results with one clear sky model, while the other cloud index gives better results with another clear sky model. It is seen that the Satel-Light clear sky model generally gives higher values than SOLIS, and that the new cloud index introduced here is generally higher than the one from Hammer (2000). Thus it is evident that sometimes errors are cancelling each other in the Heliosat-scheme, and sometimes they are adding up. Using climatological values of turbidity as input to the clear sky models, the clear sky model is by now a large uncertainty in the Heliosat-scheme. This makes it difficult to find the optimal cloud index. The inclusion of real time atmospheric parameters in the new SOLIS clear sky model in the near future looks promising. As the clear sky value will then have a stronger physical basis, both the cloud index and the relationship with the clear sky index should be suitably chosen, and perhaps 9 tuned to ground data to minimise the deviation. The remaining empirical parameters which can be adjusted for best performance are: The method to determine ρcloud. Here a 98 percentile of reflectivities is used, but there is no clear physical reasoning behind such a choice. The bias of the Heliosat-algorithm is highly sensitive to the choice of ρcloud. The method to determine ρground. The bias is not as sensitive to this parameter, as it is in both the nominator and denominator of equation 5, but it should certainly be chosen so that when the reflectivity equals ρground , the observed clear sky value is matching the clear sky model. The clear sky - cloud index relationship. For the cloud index approaching 1 the value of the clear sky index should be in accordance with the value of ρcloud. In other words; the clear sky value for the cloud index equal to 1 should be similar to what is observed under a cloud cover giving a reflectivity equal to ρcloud. It is seen that the new cloud index has a clear peak around zero, while the traditional has more frequently negative values. The reason is that the old histogram technique gives a higher value for ρground than the new algorithm presented in section 2.2. There are two ways the cloud index can become negative; one is when the atmosphere is clearer than the "reference atmosphere" for which the clear sky model applies. For this case a negative cloud index gives correctly a global irradiance which is higher than the clear sky model. A second reason can be that the pixel is completely in shadow from a nearby cloud which is not seen on this pixel. In this case the conditions are taken as very clear, while the ground measured irradiance is low due to the shadows. Probably both situations occur, and the effects are cancelling each other. That the 'traditional' cloud index is more frequently negative needs not be a bad sign, but one should be aware of the difference when it is combined with a clear sky model. For both cloud indices and clear sky models averaging the cloud index over 3x5 pixels gives best results. Acknowledgements: This work is a part of the project Heliosat-3 funded by the European Commission (NNK5-CT-20000322). We thank project colleagues for valuable discussions and advices. The following persons are thanked for providing the ground measurements for the validation: Antonio Ortegón Gallego (Barcelona), Christian Reise (Freiburg), Pierre Ineichen (Geneva) and Dominique Dumortier (Lyon). We also thank Jethro Betcke and Rolf Kuhleman at the University of Oldenburg for providing data from Meteosat-8 for specific sites. 6 References Rigollier, C; Lefèvre, M; Wald, L (2004) The method Heliosat-2 for deriving shortwave solar radiation from satellite images, Solar Energy 77, 159-169 Hammer, A (2000) Anwendungsspezifische Solarstrahlungsinformationen aus Meteosat-daten, PhD thesis, University of Oldenburg Dagestad, K-F (2001) Ein modellstudie av samanhengen mellom reflektert radians ved toppen av atmosfæra og globalstråling ved bakken, Master Thesis, University of Bergen Govaerts, Y; Clerici, M (2004) MSG-1/SEVIRI Solar Channel Calibration, ; Commisioning Activity Report for Eumetsat. EUM/MSG/TEN/04/0024, Version 1.0, 21 Jan 2004 Fontoynont, M; Dumortier, D; Heinemann, D; Hammer, A; Olseth, J A; Skartveit; Ineichen, P; Reise, C; Page, J; Roche, J; Beyer, H; Wald, L (1998) Satel-Light: A www server which provides high quality daylight and solar radiation data for western and central Europe, Proc. 9th conference on satellite meteorology and oceanography in Paris, 25-28 May 1998, pp. 434-437 10 Rigollier, C.; Wald, L. (1998) Using Meteosat images to map the solar radiation:improvements of the Heliosat method, Proceedings of the 9th Conference on Satellite Meteorology and Oceanography. Published by Eumetsat, Darmstadt,Germany, EUM P 22, pp. 432-433 Page, J (1996) Algorithms for the Satel-Light programme, Technical report for the Satel-Light programme Dumortier, D (1995) Modelling global and diffuse horizontal irradiances under cloudless skies with different turbidities, Technical report for the Daylight II project, JOU2-CT92-0144 Dumortier, D (1998) The Satel-Light model of turbidity variations in Europe, Report for the 6th Satel-Light meeting in Freiburg, Germany, September 1998 Mueller, R; Dagestad, K-F; Ineichen, P; Schroedter, M; Cros, S; Dumortier, D; Kuhlemann, R; Olseth, J. A; Piernavieja, C; Reise, C; Wald, L; and Heinemann, D (2004) Rethinking satellite-based solar irradiance modelling, The SOLIS clear sky module, Remote Sensing of the Environment 91, 160-174 Holzer-Popp, T; Schrodter, M; Gesell, G (2002a) Retrieving aerosol optical depth and type in the boundary layer over land and ocean from simultaneous GOME spectrometer and ATSR-2 radiometer measurements, 1, Method description, J. Geophys. Res. 107, AAC16-1 – AAC16-17 Holzer-Popp, T; Schrodter, M; Gesell, G (2002b) Retrieving aerosol optical depth and type in the boundary layer over land and ocean from simultaneous GOME spectrometer and ATSR-2 radiometer measurements, 2, Case study application and validation, J. Geophys. Res. Paltridge, G; Platt, C (1976) Radiative processes in meteorology and climatology, Elsevier, ISBN 0-444-41444-4 Appendix A.1 An analytical expression for the backscattered radiation from air molecules An analytical expression for the scattered radiance from the air molecules towards a satellite will here be derived for plane-parallel conditions. It is assumed that there are no other scattering or absorbing agents in the atmosphere, and that all photons are scattered only once. Figure 5 shows a situation with a solar zenith angle θ, and a satellite viewing the sunlit area at ground with a zenith angle φ. The optical depth due to scattering is increasing downwards from 0 at the top of atmosphere to τ and τ+dτ at two indicated levels. 11 Figure 5: Schematic illustration of the infinitesimal volumes of the atmosphere between the optical depth τ and τ + δτ from which solar radiation is scattered towards a satellite. θ and ф are the solar and satellite zenith angles, respectively. The amount of radiation scattered per volume unit at level τ is given by: direct radiation into volume dV −direct radiation out of dV dV I e − cos −I e −d cos d cos − −d = (8) − cos = I e cos 1−e cos ≈ I e cos d where Iλ is the incoming monochromatic irradiance at the top of the atmosphere. The scattered radiation from the volume dVφ that reaches the satellite is given by: scattered radiation × phase function×transmissivity towards satellite×volume= unit volume − − cos cos d I e P e cos (9) where P(ψ) is the scattering phase function and ψ is the angle between the directions towards the sun and satellite as seen from ground. Integration from τ = 0 at the top of the atmosphere to τ gives the scattered radiance towards the satellite from the whole atmospheric column down to the level τ: r = I P cos ∫e 0 ' − 1 1 cos cos 1 1 − cos d ' =I P [1−e cos cos ] cos cos 12 (10) The Rayleigh scattering function is well known: P = 3 2 1cos 16 (11) Note that in this context ψ is 180 degrees minus the more commonly used scattering angle, and is therefore referred to as the 'co-scattering angle'. Equation 11, however, remains unchanged. An expression for the vertical optical depth of the atmosphere due to Rayleigh scattering as a function of wavelength is found from Paltridge et al. (1976): = −4.05 0.311 m (12) A.2 Adaptation to the Meteosat-8 HRV response function Equations 10, 11 and 12 describe monochromatic radiation scattered towards a satellite. To get the actual radiance observed by Meteosat it should be integrated over all wavelengths, weighted with the Meteosat-8 HRV response function: r atm = ∫ r R HRV d ∫ RHRV d (13) Then the equation would have to be integrated numerically for any actual geometrical configuration. Figure 6 shows that by using an 'equivalent wavelength' of 680 nanometres radiances very close to what is obtained by numerical integration over all wavelengths are found. To a good approximation the following equation can therefore be used for the scattered solar radiation from air molecules reaching Meteosat-8: 1 2 1 − 31cos cos r atm =I 0 [1−e cos cos ] 16 cos cos (14) where I0 is the solar constant of 1367 W/m2. Equation 12 gives a value for the optical depth τ of 0.0426 for λ = 680 nanometres. The same equation can be used for Meteosat-7 which has the same spectral response function of the HRV-channel. For other spectral channels similar 'equivalent wavelengths' could be obtained by the same integration. It was shown in Dagestad (2001) that for the wavelengths of the Meteosat HRV function single scattering is dominant. Care should however be taken for wavelengths where multiple scattering is dominant. 13 Figure 6: Calculation of the radiance scattered from air molecules and observed by the Meteosat-8 HRV channel for various angular configurations. The x-axis is the result obtained by numerically integrating Equation 13. The y-axis is the results by using Equation 14 with optical depths for the wavelengths indicated on the figure. The geometrical configurations used are the actual configurations for all Meteosat-8 images for the period 16 March to 31 August 2004 for the same stations as in Figure 1. 14 Paper IV Dagestad, K-F. and Wald, L. (2005) Assessment of the database HelioClim-2 of hourly irradiance derived from satellite observations Draft Assessment of the database HelioClim-2 of hourly irradiance derived from satellite observations Knut-Frode Dagestad 1) and Lucien Wald 2) 1) 2) Geophysical Institute, Univeristy of Bergen, Allegaten 70, Norway Groupe Télédétection & Modélisation, Ecole des Mines de Paris / Armines, BP 207, 06904 Sophia Antipolis cedex, France Draft May 2005 Abstract This paper presents a validation of the database HelioClim-2, which provides solar irradiance data processed from satellite images on a web server (www.helioclim.net). Like HelioClim-1, HelioClim-2 is using the Heliosat-2 algorithm for the processing, but the input satellite data to HelioClim-2 are sampled at a higher frequency both in space and time. When compared to daily sums of global irradiance at five European locations for the period March to August 2004, HelioClim-2 gives a Root Mean Square Deviation (RMSD) of 16%, compared to 30% for HelioClim-1. For hourly irradiances HelioClim-2 gives an RMSD of 26%, compared to 18% for the Heliosat-1 method. It is seen that HelioClim-2 overestimates the global irradiance for all stations. A likely reason is found to be a too high value of the cloud reflectivity within the Heliosat-2 algorithm. 1 Introduction The benefit of observations made by geostationary satellite for assessing the solar irradiance –also called shortwave downwelling irradiance- has been demonstrated by a large number of studies. The Ecole des Mines de Paris has created a database called HelioClim-1 by applying the method Heliosat-2 (Rigollier et al. 2004) to images acquired by the series of Meteosat satellites, from Meteosat-4 to -7. The geographical coverage of HelioClim-1 is the field of view of Meteosat, i.e., Europe, Africa and the Atlantic Ocean. The database contains daily irradiation -or daily mean irradiance- for every day from 1985 up to present. The quality of HelioClim-1 has been assessed by several comparisons between ground measurements and HelioClim-1 values (Lefèvre et al. 2005; Cros, 2004). HelioClim-1 data are available by the means of the SoDa web service (www.sodais.com). This database has met an unexpected success: approximately 1000 requests per month in early 2005. Building on this success, Ecole des Mines de Paris took the opportunity of the launch of the new series of Meteosat satellites to create a new database HelioClim-2. This new series Meteosat Second Generation (MSG) became operational in late January 2004 and the new satellited is called 1 Meteosat-8. HelioClim-2 is also accessible by the means of the SoDa web service. It provides hourly mean irradiance from February 2004 up to present. The geographical coverage is similar to that of HelioClim-1. The present article is the first assessment of the quality of HelioClim-2, of both hourly and daily mean irradiances, with respect to ground measurements. In addition, other sources of data, e.g., HelioClim-1, are used. 2 The HelioClim-2 database The method Heliosat-2 is operated in real-time on Meteosat-8 data acquired by a receiving station at Ecole des Mines de Paris. Because Heliosat-2 has been developed to perform on images acquired in a broadband range, the MSG data acquired in the narrow visible bands VIS-1 (650 nm) and -2 (850 nm) are combined together to simulate the broadband channel of Meteosat-7 (Cros et al. 2005). The calibration coefficients are read in the header of the Meteosat-8 images on the one hand and on the Web site of Eumetsat for Meteosat-7 on the other hand (www.eumetsat.de). The radiances are remapped before they are processed by the method Heliosat-2; first they are filtered by a 17x17 filter (Aiazzi et al. 2002) to remove all fine structures and then re-mapped on a regular grid in latitude and longitude. The cell is squared and its size is 5' of arc angle, that is approximately 10 km at midlatitude. The differences between HelioClim-1 and -2 are summarized in Table 1. This table gives also an overview of other sources of data that will be discussed later. Table 1: Overview of the input data to the various implementations of the Heliosat algorithm(s) compared in this paper. An overview of the Heliosat-1 and Heliosat-2 algorithms is shown in Table 3. Satellite Sensor Type of Data Size of original pixel (nadir) Original sampling period Spatial sampling Temporal sampling Spatial averaging Algorithm used HelioClim-1 HelioClim-2 Heliosat-1, Heliosat-1, M-7 M-8 Meteosat-7 Meteosat-8 Meteosat-7 Meteosat-8 HRV HRV simulated from MSG SEVIRI-VIS-1 and -2 HRV SEVIRIHRV Reduced format IDS-B2 Re-mapped data Highresolution image Highresolution image 5 km 3 km 2.5 km 1 km 30 minutes 15 minutes 30 minutes 15 minutes every 6th pixel 5 minutes of arc ~10 km at midlatitude every pixel every pixel one image out of 6 one image out of 4 every image every image (3 hours) (one hour) none single pixel + 2 different averages 5 x 5 pixels 3 x 5 pixels Heliosat-2 Heliosat-2 (Heliosat-1) (Heliosat-1) 2 3 Ground measurements and other data for comparison Ground measurements from five sites located in Europe (Table 2) for the period 16 March to 31 August 2004 are compared to HelioClim-2 values. In addition, we take advantage of the fact that both Meteosat satellites are observing the same area for this period. Accordingly, for the same sites also HelioClim-1 values (daily values from Meteosat-7 images) are available. Also available is data made by the University of Oldenburg using the method Heliosat-1 applied to Meteosat-7 images hereafter called Heliosat-1 M-7 (see Table 1). Finally, a last series of data is available: a modified version of Heliosat-1 applied to the high spatial resolution data of Meteosat-8 - hereafter called Heliosat-1 M-8 (see Table 1). The two latter data sets are the courtesy of the University of Oldenburg. Table 3 shows the difference between Heliosat-1 and Heliosat-2. The modified Heliosat-1 uses the SOLIS clear-sky model (Mueller et al. 2004) instead of the model of diffuse irradiance from Dumortier (1995) and direct irradiance from Page (1996). Only days for which data are available from all stations and all four models for all hours were used, ensuring that all models are compared with the exact same measurements. This included 155 days with 1475 hourly values. Table 2: Global irradiances from these five stations are compared to satellite derived data. Elevation Latitude Longitude Station [m] [ºN] [ºE] Instrument Barcelona 98 41.39 2.12 Kipp & Zonen CM 11 Bergen 45 60.40 5.32 Kipp & Zonen CM 11 Freiburg 275 48.02 7.84 Kipp & Zonen CM 11 Geneva 425 46.20 6.13 Kipp & Zonen CM 10 Lyon 170 45.78 4.93 Kipp & Zonen CM 6 Table 3: Differences between the versions of the Heliosat algorithm used in this paper. Heliosat-1 is the version described in Fontoynont et al. (1998) and Heliosat-2 is described in Rigollier et al. (2004). Heliosat-1 Heliosat-2 Input data from satellite raw counts calibrated radiances Backscatter correction Empirical relationship from Hammer (2000) Ground albedo Histogram technique from Hammer (2000) Based on the diffuse ESRA clear sky model, (Rigollier, 2000) Second minima of time series of reflectivites Cloud albedo Constant of 150 normalized satellite counts Empirical model based on Taylor & Stowe (1984) Clear sky model Model of diffuse irradiance from Dumortier (1995) and direct irradiance from Page (1996). Input is Linke turbidities from Dumortier (1998) ESRA clear sky model, Rigollier (2000) modified Geiger et al. Input is Linke turbidities from Lefevre et al. 2004 cloud index - clear sky index relationship Empirical relationship from Rigollier et al. (1998) Empirical relationship from Rigollier et al. (1998) 3 4 Comparison with ground data 4.1 Daily values Table 4 shows a comparison of daily means of solar irradiance from HelioClim-1 and HelioClim-2 versus observations from the five European stations in Table 2. It is seen that HelioClim-2 gives much better results than HelioClim-1, except for the station in Bergen. The average Root Mean Square Deviation (RMSD) for HelioClim-2 is 16%, compared to 30% for HelioClim-1. However, for HelioClim-2 there is a positive bias for all stations, and for Bergen it is as large as 23%. For HelioClim-1 there is a large bias of 25% for Barcelona. The reasons for these biases will be discussed in section 5. Table 4: Root Mean Square Deviation (RMSD) and Mean Bias Deviation (MBD, model - observation) for daily means of global irradiance for HelioClim-1 and HelioClim-2. Values are given in W/m2 with percentages of the mean observed values in parantheses. Observed HelioClim-1 HelioClim-2 [Wm-2] MBD RMSD MBD RMSD 211 53 (25) 78 (37) 20 (10) 30 (14) Bergen 97 8 ( 6) 36 (27) 31 (23) 41 (31) Freiburg 172 -4 (-3) 41 (24) 2 ( 1) 22 (13) Geneva 196 -12 (-6) 52 (27) 7 ( 3) 22 (11) Lyon 183 2 ( 1) 49 (27) 8 ( 4) 18 (10) All stations 179 9 ( 5) 52 (30) 13 ( 8) 28 (16) Station Barcelona 4.2 Hourly values As HelioClim-1 does not provide hourly values, HelioClim-2 is compared to Heliosat-1 using Meteosat-7 data. While Heliosat-1 uses full resolution satellite data as input, the input to HelioClim-2 are sampled in both time and space; see Table 1 for an overview of the differences. Table 5 shows a comparison of hourly global irradiances from HelioClim-2 and Heliosat-1 versus observations. In average, Heliosat-1 is seen to give the smallest RMSD, with 18% compared to 26% for HelioClim-2. In section 5 it will be assessed how much of this difference is due to the spatial and temporal samling of satellite data, and how much is due to differences of the Heliosat-1 and Heliosat-2 algorithms. 4 Table 5: Root Mean Square Deviation (RMSD) and Mean Bias Deviation (MBD, model - observation) for hourly global irradiance for Heliosat-1 and HelioClim-2. Values are given in W/m2 with percentages of the mean observed values in parantheses. Station Observed Heliosat-1, M-7 HelioClim-2 -2 [Wm ] MBD RMSD MBD RMSD Barcelona 531 2 ( 0) 85 (16) 52 (10) 114 (22) Bergen 280 26 ( 9) 82 (29) 59 (21) 112 (40) Freiburg 389 -5 (-1) 72 (19) 2 ( 1) 102 (26) Geneva 459 2 ( 0) 73 (16) 16 ( 3) 103 (23) Lyon 431 7 ( 2) 63 (15) 24 ( 6) 91 (21) All stations 409 7 ( 2) 75 (18) 31 ( 8) 105 (26) 5 Analysis of the results Section 4 shows that HelioClim-2 overestimated the global irradiance for all the five stations, and particularly for Bergen with a bias larger than 20%. The frequency distributions of hourly clearness indices (global irradiance divided by irradiance at the top of the atmosphere) on Figure 1 suggest an explanation for this: for all stations HelioClim-2 gives too few clearness indices below 0.2. Hence HelioClim-2 overestimates the global irradiance for cloudy conditions. Since Bergen is the site with the most frequent cloudy conditions and also the site with the largest bias, it is likely that the reason for the overall overestimation by HelioClim-2 is that it gives too low cloud index for the cloudy cases, and thus too high irradiance. 5 Figure 1: Frequency distributions of clearness indices for HelioClim-2 (dashed line) and the observations (solid line) for the five stations in Table 5. The cloud index n is given by: n= − ground cloud − ground (1) where the parameters ρground and ρcloud are the reflectivities of the ground and the thickest clouds, respectively. A too low cloud index for the cloudy cases (ρ ≈ ρcloud ) will occur when ρcloud is too high. The effect of a too high value of ρcloud can be illustrated with a case study: A simplified version of Heliosat is applied to the Meteosat-8 data from Bergen, where constant values have been used for the parameters ρground and ρcloud. The solid line on Figure 2 shows the observed clearness indices for Bergen, and the solid line with circles is a 'best fit' to the observations with ρground and ρcloud tuned to 0.16 and 0.77, respectively. For this case the RMSD and MBD is 23% and 0%, respectively. With ρcloud increased to 0.95 (dashed line) the most frequent 'low clearness index' is increased from approximately 0.15 to 0.30, like for the HelioClim-2 values (Figure 1, 'Bergen'). The RMSD is now 31% and the MBD is 16%. For this case study the reflectivities were calculated from 3x5 averages of Meteosat-8 counts with backscatter correction from Dagestad (2005). 6 Figure 2: Frequency distributions of hourly clearness indices for Bergen. The solid line is the observations and the solid line with circles is a simplified version of Heliosat with a constant 'cloud albedo' of 0.77. For the dashed line the cloud albedo is increased to 0.95. See the text for details. The above example suggests that the 'apparent cloud albedo' in the HelioClim-2 algorithm is too high by an absolute value of approximately 0.18. From Rigollier et al. (2004) ρcloud is given by: cloud = eff −atm T s T v (2) ρeff is here an effective cloud albedo (cloud + atmosphere) from Nimbus-7 measurements given by Taylor & Stowe (1984). ρatm is in principle the radiance scattered by the atmospheric layer above clouds, but is modelled by the path radiance for the whole atmosphere. T(θs) and T(θv) are the transmissivities downwards from the sun and upwards towards the satellite, respectively, where θS and θV are the solar and satellite zenith angles, respectively. These two transmissivity factors might be a cause of a too high apparent cloud albedo: the expression used is from the ESRA model (Rigollier 2000), and is the transmissivity down to ground level. The transmissivity down to the level of the cloud tops should however be larger, and hence equation 2 should give too high value, subsequently giving too low cloud index and then too high clear sky index. In addition to a too high cloud albedo, the sampling of satellite data may introduce a bias; Tables 4 and 5 show that both HelioClim-1 and HelioClim-2 overestimates irradiance for Barcelona and Bergen. These two sites are located by the coast, and the spatial sampling and filtering of satellite 7 data may therefore include measurements over sea. Since there are generally less clouds over sea than over land this can give a cloud index which is lower than what whould have been measured for the pixel covering the measuring station on land. << This paper is a draft, and Lucien Wald will later add a discussion in sections 5.1 and 5.2 about the effect of sampling and averaging the satellite data. Section 6 gives by now preliminary conclusions >> 5.1 Influence of the averaging by comparison of ground, HC2 single pixel, HC2 average and HC2 weighted average Table 6.Root Mean Square Deviations (RMSD) for hourly and daily irradiances for HelioClim-2 for three different averges. 'Simple average' means averaging 5 pixels (North, South, East, West and the central pixel), for the 'weighted average' the central pixel has got a weigth of 2. Units are W/m2 , with percentages of the mean observed irradiances in parantheses. The observed hourly and daily mean values are given in tables 4 and 5 respectively. Hourly Station Single pixel Daily Simple Weighted average average Single pixel Simple Weighted average average Barcelona 114 (22.0) 122 (23.0) 123 (23.2) 721 (14.3) 793 (15.7) 810 (16.0) Bergen 112 (40.0) 104 (37.2) 107 (38.0) 1005 (31.1) 913 (28.2) 942 (29.1) Freiburg 102 (26.3) 97 (25.0) 521 (12.7) 472 (11.5) 490 (11.9) Geneva 103 (22.6) 100 (21.7) 100 (21.7) 534 (11.4) 565 (12.0) 539 (11.5) Lyon 91 (21.1) All stations 95 (24.5) 88 (20.3) 87 (20.2) 425 ( 9.7) 418 ( 9.5) 406 ( 9.2) 105 (25.5) 102 (24.8) 103 (25.0) 673 (15.7) 660 (15.4) 669 (15.6) 5.2 Influence of the spatial sampling using Oldenburg MSG (1 km) and ground data Table 7: Root Mean Square Deviations (RMSD) of the Heliosat-1 method using Meteosat-7 and Meteosat-8 data as input. Percentages of the mean observed irradiances are given in parantheses. The last column shows the improvement in percent by using Meteosat-8 HRV as input to Heliosat-1 instead of Meteosat-7 HRV. Station Observed Heliosat-1, M-7 Heliosat-1, M-8 Improvement [Wm-2] Barcelona 531 85 (16) 72 (14) 15 Bergen 280 82 (29) 65 (23) 20 Freiburg 389 72 (19) 64 (17) 11 Geneva 459 73 (16) 64 (14) 13 Lyon 431 63 (15) 58 (13) 9 All stations 409 75 (18) 64 (16) 14 8 6 Conclusions Global irradiances from the database HelioClim-2 have been compared with ground measurements for five European stations. The results are compared to HelioClim-1 for daily values and to Heliosat-1 for hourly values. For the daily values HelioClim-2 is a large improvement with a Root Mean Square Deviation (RMSD) of 16%, compared to 30% for HelioClim-1. Since both databases are based on the same algorithm, Heliosat-2, the improvement must be due to the increased spatial and temporal sampling of the input data to HelioClim-2 (Table 1). Hourly global irradiances from HelioClim-2 are compared with ground measurements and with output from the Heliosat-1 algorithm. The mean RMSD for Heliosat-1 is 18%, and for HelioClim-2 it is 26%. HelioClim-2 is seen to overestimate the global irradiance for all stations, and particularly for Bergen where the bias is 21%. It is pointed out that a likely explanation is a too high value of the parameter ρcloud, which in the Heliosat algorithm is the reflectivity of the thickest clouds. A reason for this can be that the reflectivity is normalised with the transmissivity down to ground level, whereas the transmissivity down to the cloud tops should have been used instead. When ρcloud is too high the cloud index will be too low, and hence the estimated irradiance will be too large. Cros S., Albuisson M., Wald L., (2005) Simulating Meteosat-7 broadband radiances at high temporal resolution using two visible channels of Meteosat-8 Dagestad, Knut-Frode, (2005) A new algorithm for calculating the cloud index. Dumortier, D,(1998), The Satel-Light model of turbidity variations in Europe, Dumortier, D,(1995), Modelling global and diffuse horizontal irradiances under cloudless skies with different turbidities, Fontoynont, M; Dumortier, D; Heinemann, D; Hammer, A; Olseth, J A; Skartveit; Ineichen, P; Reise, C; Page, J; Roche, J; Beyer, H; Wald, L, (1998) Satel-Light: A www server which provides high quality daylight and solar radiation data for western and central Europe, Proc. 9th conference on satellite meteorology and oceanography in Paris, 25-28 May 1998, pp. 434-437 Hammer, A, (2000) Anwendungsspezifische Solarstrahlungsinformationen aus Meteosat-daten, PhD thesis, University of Oldenburg Page, J,(1996), Algorithms for the Satel-Light programme, Rigollier C; Bauer O; Wald L;, (2000) On the clear sky model of the 4th European Solar Radiation Atlas with respect to the Heliosat method, Solar Energy, 68(1), 33-48 Rigollier, C; Lefèvre, M; Wald, L, (2004) The method Heliosat-2 for deriving shortwave solar radiation from satellite images, Solar Energy, 77, 159-169 Rigollier, C.; Wald, L., (1998) Using Meteosat images to map the solar radiation:improvements of the Heliosat method, Proceedings of the 9th Conference onSatellite Meteorology and Oceanography. Published by Eumetsat, Darmstadt,Germany, EUM P 22, pp. 432-433 Taylor, V.R.; Stowe, L.L.,(1984), Atlas of reflectance patterns for uniform Earth and cloud surfaces (Nimbus 7 ERB - 61 days), 10, july 1984, Washington, DC, USA Aiazzi B., Alparone L., Baronti S., Garzelli A., Context-driven fusion of high spatial and spectral resolution images based on oversampled multiresolution analysis. IEEE Transactions on Geosciences and Remote Sensing, 40, 2300-2312, 2002. Cros S., 2004, Création d’une climatologie du rayonnement solaire incident en ondes courtes à l’aide d’images satellitales (Design of an incident shortwave solar radiation climatology using satellite images). Thèse de Doctorat en Energétique, Ecole des Mines de Paris, 13 septembre 2004, 157 pages. Cros S., Albuisson M., Wald L., Simulating Meteosat-7 broadband radiances at high temporal resolution using two visible channels of Meteosat-8. To be published by Solar Energy, 2005. Geiger M., Diabaté L., Ménard L., Wald L., 2002. A web service for controlling the quality of 9 measurements of global solar irradiation. Solar Energy, Vol. 73, No 6, pp. 475-480. Lefèvre M., Albuisson M., Ranchin T., Wald L., Remund J., 2004. Fusing ground measurements and satellite-derived products for the construction of climatological maps in atmosphere optics. In Proceedings of the 23rd EARSeL Annual Symposium "Remote Sensing in Transition", 2-4 June 2003, Ghent, Belgium, Rudi Goossens editor, Milpress, Rotterdam, Netherlands, pp. 85-91. Lefèvre M., Diabaté L., Wald L., Using reduced satellite data sets ISCCP-B2 to assess surface surface solar irradiance. Submitted to Solar Energy, 2005. Mueller R.W, Dagestad K.F, Ineichen P, Schroedter M, Cros S, Dumortier D, Kuhlemann R, Olseth J.A, Piernavieja G, Reise C, Wald L, Heinnemann D, Rethinking satellite based solar irradiance modelling - The SOLIS clear sky module. Remote Sensing of Environment, 91, 160-174, 2004. 10 Paper V Dagestad, K-F. (2005) Simulations of bidirectional reflectance of clouds with a 3D radiative transfer model Manuscript Simulations of bidirectional reflectance of clouds with a 3D radiative transfer model Knut-Frode Dagestad Geophysical institute University of Bergen, Norway May 2005 Abstract To get an impression of the impact of the hetereogeneity of cloud properties on the radiance observed by a satellite for various pixel sizes, a 3-dimensional radiative transfer model (SHDOM) is used to simulate radiances reflected to space from two cloud fields. For a given sun-cloud-satellite geometry and for a typical stratocumulus cloud field the radiance is mainly unaffected by rotating the cloud in the azimuth direction when the pixel size is 3.5 kilometres. For a pixel size of 550 metres there is some more variability, and for a pixel of 55 metres the variability of the radiance is larger than 100%. As expected, the variability is much larger for a typical cumulus cloud than for the stratocumulus field. Even for a field as large as 6.7 kilometres there is some variability in the radiance by rotating the cloud field. For pixels of 670 metres or smaller the variability is extreme, as clouds are obstructing the viewing path for some rotations and not for others. The angular distribution of simulated radiances from the stratocumulus field is also compared to reflectances measured with the Meteosat High Resolution Visible sensor. The angular distribution of the simulated radiances is similar to what is observed by Meteosat for thin clouds. For the thicker clouds the distribution of Meteosat reflectances is closer to lambertian. 1 Introduction For many purposes it is useful to have a description of the fraction of solar radiation that clouds reflect in different directions. However, a general function can never be found, as the variable 3dimensional structure of clouds makes the reflectance literally unpredictable. The Heliosat-algorithm (e.g. Cano et al. 1986, Beyer et al. 1996, Fontoynont et al. 1998, Rigollier et al. 2004) estimates global irradiance from satellite images by a two step process: first, the combined reflectance from both ground and clouds, measured by the high resolution visible sensor, is used to calculate the "cloud index", a single parameter describing the cloud cover. second, this cloud index is combined with a clear sky model to estimate the actual global irradiance at ground. This study addresses one of the implicit assumptions in this approach: that a cloud field is uniquely determined by its reflectance in one particular direction. A 3-dimensional radiative transfer model, SHDOM, will be used to assess the variability related to this assumption. SHDOM is used to simulate the reflected radiances in different directions from two different cloud fields. One of the cloud fields is a rather homogenous stratocumulus field, while the second is a broken cumulus field. Keeping the sun-ground-satellite geometry (and everything else) constant, the cloud fields are rotated 0, 90, 180, and 270 degrees in the azimuth direction. So, with the cloud properties constant (except for the rotation) the variability of the reflectance, and hence the cloud index, will be investigated. Finally the spatial distribution of reflectance will be compared with actual reflectances (though normalised) from the High Resolution Visible sensor of the Meteosat satellite. 1 2 The 3D radiative transfer model SHDOM SHDOM is an acronym for "The Spherical Harmonics Discrete Ordinate Method" for threedimensional atmospheric radiative transfer, and it is developed by Frank Evans (Evans 1998). The code combines both discrete ordinates and spherical harmonics to solve the radiative transfer equation. Spherical harmonics are used to compute the source functions and the scattering integral. This method saves a lot of computer memory, since in practice the source function is often zero or smooth for large parts of the medium, and hence can be represented with few spherical harmonic terms. Another advantage is that the scattering integral is more efficiently computed in spherical harmonics than in discrete ordinates. Discrete ordinates are used to compute the radiance field, which is then used to compute the source function, and the process repeats until a stable solution is found. To speed up calculations and to save memory an adaptive grid is used; i.e. the model can start with a rather coarse grid, and fills inn extra grid points for better accuracy whenever gradients exceed a certain threshold. 3D radiative transfer calculations consume a lot of computer memory compared to 1D calculations, so the methods used to save memory makes it possible to have cloud fields with adequate spatial resolution even with 1 GB of memory. The model is not restricted to only atmospheric calculations; the input medium can be specified completely generally with extinction, single scattering albedo, scattering phase function and temperature. The simulations are done for non-polarized and monochromatic radiation, but the correlated k-distribution method can subsequently be used for integration over a spectral band. The horizontal boundaries of the input/output field may be specified as open or periodic, and the latter is used for the simulations in this paper. 3 Input data Two cloud fields are used as input to SHDOM: An overcast stratocumulus field obtained from a Large Eddy Simulation from the 1987 FIRE experiment (Moeng et al. 1996) A broken cumulus cloud field, reconstructed from a Landsat image by Frank Evans, the author of SHDOM Both cloud fields were used in the second round of the Intercomparison of 3D Radiation Codes, I3RC (http://i3rc.gsfc.nasa.gov/cases_new.html). Each grid point of the cloud fields contains a droplet effective radius in micrometres and a Liquid Water Content (LWC) in grams of liquid water per cubic metres. For the runs with SHDOM a gamma distribution with a shape parameter of 7.5 was used for the droplet size distribution. The cloud fields were then combined with a Mie scattering table for the wavelength of 670 nanometres to specify the extinction, single scattering albedo and phase function at each grid point. Table 1 shows technical data of the two cloud fields, and Figures 1a and 1b and show the vertical Liquid Water Path (LWP). There were no aerosols in the atmosphere, and the ground albedo was set to zero. Table 1: Technical data about the two cloud fields input to SHDOM Stratocumulus Cumulus Horizontal pixel size [m] 55 67 Number of pixels horizontally 64 100 Number of pixels vertically 16 36 3520 6700 Total horizontal size [m] 2 Figure 1a: Vertically integrated liquid water content (g/m2) for the stratocumulus cloud field. Figure 1b: Vertically integrated liquid water content (g/m2) for the cumulus cloud field. 3 4 Results Two slightly different experiments are performed: In section 4.1 the variability of the reflectances from two cloud fields is investigated by rotating the cloud fields while keeping everything else constant. This will be done for different sizes of the observed area of the cloud field. Thus, this will show to what extent the 3-dimensional inhomogeneities of the clouds affect radiances observed by a satellite sensor, and the pixel size needed to minimise the variability. In section 4.2 the variation of the bidirectional reflectance with the sun-cloud-satellite geometry is analysed. Here, simulations with an overcast stratocumulus cloud field are compared with reflectances observed with Meteosat. 4.1 Variability of simulated radiances 4.1.1 Experimental setup 3D simulations consume much more computer power than 1-dimensional models, so only some case studies were performed with SHDOM. The simulated radiances are monochromatic at the wavelength 670 nanometres. For each cloud field upward radiances were calculated for the following configurations: The cloud field was rotated 0, 90, 180 and 270 degrees in the azimuth direction Six different solar zenith angles were used: 0, 30, 60, 70, 80 and 85 degrees For each rotation of the cloud field and each solar zenith angle, radiances were calculated for three different viewing (satellite) zenith angles: 45, 70 and 80 degrees. The view azimuth angle was always the same as the solar azimuth angle. Radiances were averaged over three different subsets of the actual cloud fields: - The pixel in the middle of the cloud field - 10x10 pixels in the middle of the cloud field - The whole cloud field This corresponds to "satellite pixel sizes" of 55, 550 and 3520 metres, respectively, for the stratocumulus field and 67, 670 and 6700 metres, respectively, for the cumulus field (Table 1). The experiment is performed to quantify the variation of the observed radiance for different "satellite pixels sizes" when the cloud field is rotated. Since the cloud fields remain unchanged the rotation should then isolate the effect of heterogeneity from other cloud properties. 4.1.2 The stratocumulus cloud field The upper part of Figure 2 shows the simulated radiances from the stratocumulus cloud field with a view zenith angle of 45 degrees. It is seen that the radiance integrated over the whole cloud field is practically independent of the rotation of the cloud field. A noticeable variability of the radiance is seen for the 10 times 10 pixel subset and for the single pixel the radiance is varying by more than 100 percent. For all subsets a significant decrease of the reflectance is seen when the solar zenith angle is larger than 60 degrees. An explanation for this could be that the "bumps" on the top of the cloud field is scattering a large fraction of the radiation close to the forward direction, which is characteristic of Mie scattering. Thus a large part of this radiance can escape the cloud field with only a single or few scattering events. The results for a view zenith angle of 70 degrees are seen on the lower part of Figure 2. The 4 radiance from the whole cloud field is still unaffected by the rotation of the cloud field. For the smaller subsets there is a somewhat smaller variability than for a view zenith angle of 45 degrees, both in relative and absolute differences. An exception is for solar zenith angles larger than 60 degrees, where the variability for the subsets is similar for both view zenith angles. For a view zenith angle of 70 degrees a peak in the reflectance is seen when the solar zenith angle is close to the view zenith angle. This is the opposition effect: no shadows are seen when the sun and the observer is in the same direction, and the cloud looks brighter. This effect is not clearly seen for the view zenith angle of 45 degrees (upper part of Figure 2). The results from a view zenith angle of 80 degrees are very similar to the case with view zenith angle of 70 degrees, and the results are therefore not shown. Figure 2: Simulated radiances from the stratocumulus cloud field for view zenith angles of 45 degrees (upper part) and 70 degrees (lower part) plotted versus the solar zenith angle. The left figures show the mean radiance from a single pixel in the middle of the cloud fields, the middle figures are for a subset of 10x10 pixels in the middle of the cloud field, and the rightmost figures show the mean radiance from the whole cloud field (64x64 pixels). The different lines are the radiances after rotating the cloud field 0, 90, 180 and 270 degrees in the azimuth direction. The incoming irradiance on a horizontal plane at the top of the atmosphere is one unit. The view azimuth angle is always the same as the solar azimuth angle. Note different scale on y-axes for upper and lower part. 4.1.3 The cumulus cloud field Radiances from the cumulus cloud field with a view zenith angle of 45 degrees are shown on the upper part of Figure 3. As expected, the variability is larger than for the stratocumulus field. Even the mean radiance from the whole cloud field (6700x6700 metres) shows some, although not dramatic, variability from the rotation. However, for the smaller subsets, the variability is very large, with observed radiances close to zero for some of the rotations. Although the same pixels at 5 the top of the cloud field are viewed from above, the path to the ground is obstructed by clouds at a lower level for to of the rotations, while the path is clear for the others. For a view zenith angle of 70 degrees the results are similar (lower part of Figure 3). But for the cumulus cloud field no clear opposition effect like in Figure 2 is observed. The reason is probably that there are very little shadows on the scattered clouds in the cumulus field. The large decrease of radiance for solar zenith angle above 60 degrees, which is observed for the stratocumulus field, is not seen for the cumulus field. Since the cumulus clouds are thicker than the "bumps" on top of the stratocumulus field, radiation penetrating the clouds probably encounters multiple scattering events, also when the radiation is coming from a very high solar zenith angle. Thus, fewer photons escape in the forward direction for low sun with the cumulus field than with the stratocumulus field. Like for the stratocumulus cloud field the results for a view zenith angle of 80 degrees are similar to those for a view zenith angle of 70 degrees, and hence these results are not shown. Figure 3: Simulated radiances from the cumulus cloud field for view zenith angles of 45 degrees (upper part) and 70 degrees (lower part) plotted versus the solar zenith angle. The left figures show the mean radiance from a single pixel in the middle of the cloud fields, the middle figures are for a subset of 10x10 pixels in the middle of the cloud field, and the rightmost figures show the mean radiance from the whole cloud field (100x100 pixels). The different lines are the radiances after rotating the cloud field 0, 90, 180 and 270 degrees in the azimuth direction. The incoming irradiance on a horizontal plane at the top of the atmosphere is one unit. The view azimuth angle is always the same as the solar azimuth angle. Note different scale on y-axes for upper and lower part. 6 4.2 Comparison with bidirectional reflectance from Meteosat In this section the reflected radiances simulated with SHDOM are compared to measurements from the High Resolution Visible (HRV) sensor of the Meteosat satellite. A description of the full 3dimensional distribution of reflectances is difficult, so a single angular parameter will be used for this purpose. The angle between the directions towards the sun and the satellite as seen from the cloud ("co-scattering angle", ψ) is convenient; the scattering phase function of cloud droplets depend solely on this angle, and besides this is the single parameter best describing the amount of shadows seen on a rough surface. 4.2.1 Bidirectional reflectance observed with the Meteosat HRV sensor Pixel counts from Meteosat-5 for all images in 1996 have been extracted for the 61 locations shown on Figure 4. This amounts in total to 425711 values. The HRV channel of Meteosat has a spectral response function from 0.45 to 1.0 micrometres and so responds to visible and near infrared radiation. The raw pixel counts, subtracted the constant offset of 5 digital counts, are normalized with the incoming irradiance at the top of the atmosphere. Scattered radiation from atmospheric molecules is then subtracted with an empirical expression from Hammer (2000) and the values are then put into bins for each 5th degree of the co-scattering angle ψ. Within each bin the 5, 15, 35, 55, 75, 95 and 98 percentiles are calculated. These percentiles are intentionally representing increasing cloud thickness and/or amounts, provided that the cloud properties are independent of the sunground-satellite geometry. The 5 percentile is probably for cloud free cases, and is therefore representative of the ground reflectance. Since the Meteosat data are not calibrated, only the relative variation with geometry will be compared with the corresponding variation of the SHDOM results. Therefore the reflectances are divided by the value of the bin 0º < ψ < 5º for each percentile. These normalised reflectances are plotted versus ψ on Figure 5 (solid lines). Figure 4: A Meteosat image from Europe showing the location of the 61 pixels for which the variation of bidirectional reflectance has been investigated. 7 4.2.2 Bidirectional reflectance simulated with SHDOM Only the stratocumulus cloud field (Figure 1a) is used for this section. The setup of the simulations is identical to the description in section 4.1.1, but more view angles are used: For each of the solar zenith angles, radiance is calculated for every 5th degree of the view zenith angle from 0 to 85 degrees and every 15th degree of the view azimuth angle from 0 to 360 degrees. The cloud field was not rotated in the azimuth direction, and only the mean radiance from the whole cloud field was used. Since the incoming irradiance at the top of the cloud field is one unit on a horizontal plane, the radiances are equivalent to bidirectional reflectances. The solid line with the black dots on Figure 5 is the mean radiance calculated within the same bins of the co-scattering angle as for the Meteosat-data. Like for the Meteosat data, the simulated reflectances have also been normalised with the mean value for ψ less than 5 degrees, so that only the relative variation of reflectance with ψ is shown. 4.2.3 Comparison of the bidirectional reflectances Figure 5 shows that the various percentiles of the normalised bidirectional reflectances from Meteosat (solid lines) have a local maximum for ψ below 5 degrees. Then they decrease with ψ until approximately 70 degrees, from where the reflectance is increasing. The variation with ψ is largest for cases with little or no clouds; for thicker clouds (75 and 95 percentiles) the reflectance is close to lambertian. The 98 percentiles of the reflectances are, however, increasing very much when ψ is larger than 30 degrees. Hence in some rare cases the reflectance measured from Meteosat can deviate strongly from lambertian distribution also for very thick clouds. The normalised radiances from SHDOM (solid line with black dots on Figure 5) have a variation with ψ which is similar to the 15-percentile of the Meteosat-reflectances. However, the simulated radiances increase faster with ψ than the Meteosat-reflectances when ψ is larger than 70 degrees. One of the purposes of this study was to fit a one-parameter (ψ) function to the SHDOM-radiances to be used as a Bidirectional Reflectance Distribution Function (BRDF) for the maximum cloud reflectance in the Heliosat-scheme. However, it appears that the stratocumulus cloud field is too thin to represent the variation of reflectance with the geometry of a thick cloud cover. From the percentile-plots of the Meteosat-reflectances shown on Figure 5 it is also seen that the angular variation of reflectance is larger for low and intermediate cloudiness than for the thickest clouds. 8 Figure 5: Relative variation of the reflectance of clouds versus the co-scattering angle ψ. The line with dots is the reflectance of a stratocumulus cloud field (Figure 1) calculated with SHDOM (section 4.2.2). The lines without dots are various percentiles of the reflectances measured with the Meteosat HRV sensor for 1996 (section 4.2.1). Each line is labelled with the corresponding percentile number. Both the simulated and observed reflectances are normalised with the corresponding mean value for ψ less than 5 degrees. 5 Summary and conclusions The variability of bidirectional reflectance of clouds due to 3-dimensional inhomogeneities has been investigated with the 3-dimensional radiative transfer model SHDOM. For very small subsets (~50 metres) of larger cloud fields, there is significant variability of reflectance when the cloud fields are rotated in the azimuth direction. For larger pixel sizes there is a smoothing due to bright and shaded parts cancelling each other. For an overcast stratocumulus cloud field there is almost no variation in the reflectance due to the rotation for a cloud field of the size of 3520 metres. For a broken cumulus field there is some variability (~10-20 %) even for a cloud field at the size of 6700 metres. Hence, for the Heliosat method, averaging reflectances over an area of ~10 kilometres will avoid most variability due to cloud inhomogeneities. The angular variation of bidirectional reflectance from clouds has also been investigated using both simulations with the stratocumulus cloud field and measurements from the High Resolution Visible channel of the Meteosat satellite. It is found that the simulated reflectances show large variations with the angle between the directions towards the sun and satellite (ψ). The variation is similar to the variation for thin clouds observed by Meteosat. For the thicker clouds, the Meteosat-reflectances are close to lambertian. Hence, the assumption in the Heliosat method of lambertian reflectance from the thickest clouds is reasonable. However, the reflectance varies more with ψ for intermediate cloudiness, so if possible a correction should be made for this in the future. 9 References Cano D; Monget JM;, Albuisson M; Guillard H; Regas N; Wald L; (1986) A method for the determination of the global solar radiation from meteorological satellite data, Solar Energy, 37, 3139 Beyer H.G.; Costanzo C.; Heinemann D; (1996) Modifications of the heliosat procedure for irradiance estimates from satellite images, Solar Energy, 56, 207-212 Fontoynont, M; Dumortier, D; Heinemann, D; Hammer, A; Olseth, J A; Skartveit; Ineichen, P; Reise, C; Page, J; Roche, J; Beyer, H; Wald, L (1998) Satel-Light: A www server which provides high quality daylight and solar radiation data for western and central Europe, Rigollier, C; Lefèvre, M; Wald, L (2004) The method Heliosat-2 for deriving shortwave solar radiation from satellite images, Solar Energy, 77, 159-169 Evans, K. F (1998) The spherical harmonic discrete ordinate method for three-dimensional atmospheric radiative transfer, J. Atmos. Sci., 55, 429-446 Moeng, C.-H., W. R. Cotton, C. Bretherton, A. Chlond, M. Khairoutdinov,S. Krueger, W. S. Lewellen, M. K. MacVean, J. R.M. Pasquier, H. A. Rand, A. P. Siebesma, B. Stevens, and R. I.Sykes, (1996) Simulations of a stratocumulus-topped planetary boundary layer: Intercomparison among different numerical codes, Bull. Amer. Meteor. Soc., 40, 51-69 Hammer, A (2000) Anwendungsspezifische Solarstrahlungsinformationen aus Meteosat-daten, PhD thesis, University of Oldenburg 10
© Copyright 2025