13.12.2012 Views

Evaluation of a probabilistic cloud masking algorithm for ... - CM SAF

Evaluation of a probabilistic cloud masking algorithm for ... - CM SAF

Evaluation of a probabilistic cloud masking algorithm for ... - CM SAF

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

EUMETSAT Satellite Application Facility on Climate Monitoring<br />

Visiting Scientist Report<br />

<strong>Evaluation</strong> <strong>of</strong> a <strong>probabilistic</strong> <strong>cloud</strong> <strong>masking</strong><br />

<strong>algorithm</strong> <strong>for</strong> climate data record processing:<br />

SPARC: a new scene identification<br />

<strong>algorithm</strong> <strong>for</strong> MSG SEVIRI<br />

CDOP VS Study No 14<br />

Date 15. October 2010<br />

Reference CLM_VS10_03


SPARC: a new scene identification<br />

<strong>algorithm</strong> <strong>for</strong> MSG SEVIRI<br />

Report on Modifications and Validation<br />

Fabio Fontana<br />

University <strong>of</strong> Bern<br />

Reto Stockli<br />

MeteoSwiss<br />

Stefan Wunderle<br />

University <strong>of</strong> Bern<br />

September 2010


Summary<br />

The Separation <strong>of</strong> Pixels Using Aggregated Rating Over Canada (SPARC) algo-<br />

rithm represents an alternative to conventional scene identification <strong>algorithm</strong>s as<br />

it outputs a single <strong>cloud</strong> contamination rating (F) instead <strong>of</strong> a <strong>cloud</strong> mask with<br />

a limited number <strong>of</strong> categories. The rating itself is mainly calculated from three<br />

sub-scores generated based on in<strong>for</strong>mation on the brightness temperature (T-score),<br />

brightness (B-score), and the reflectance in the visible and short-wave infrared por-<br />

tion <strong>of</strong> the electromagnetic spectrum (R-score). Even though the <strong>algorithm</strong> was<br />

originally designed <strong>for</strong> the Advanced Very High Resolution Radiometer (AVHRR),<br />

it may be applied to data from other multispectral sensors.<br />

This report describes a modified version <strong>of</strong> SPARC and its application to data from<br />

the Meteosat Second Generation (MSG) Spinning Enhanced Visible and Infrared<br />

Imager (SEVIRI) sensor. A validation was per<strong>for</strong>med <strong>for</strong> the period from July<br />

2004 to June 2005. SPAR<strong>CM</strong>SG output was compared to <strong>cloud</strong> in<strong>for</strong>mation from<br />

the Alpine Surface Radiation Budget (ASRB) Network in the Swiss Alps: Cloud<br />

Free Index (CFI) and Partial Cloud Amount (PCA; in eighths). Results show a<br />

good agreement between CFI and F mainly <strong>for</strong> daytime observations, with linear<br />

correlation coefficients (r) ranging between 0.67 in fall and 0.78 in the winter months.<br />

At the ASRB site in Payerne, observations <strong>of</strong> complete clear-sky agreed with 85%<br />

(92.3%) during daytime (nighttime) between ASRB PCA and SPAR<strong>CM</strong>SG <strong>cloud</strong><br />

mask. However, results also point to an underestimation <strong>of</strong> nighttime <strong>cloud</strong> cover<br />

by SPAR<strong>CM</strong>SG. If only two SEVIRI spectral channels are fed into SPARC to simulate<br />

data from the Meteosat First Generation (MFG), a good agreement between CFI and<br />

F is obtained during daytime (r=0.65). Finally, application <strong>of</strong> SPARC at full disk<br />

level revealed an underestimation <strong>of</strong> <strong>cloud</strong> cover compared to the Climate Monitoring<br />

(<strong>CM</strong>)-Satellite Application Facility (<strong>SAF</strong>) <strong>cloud</strong> mask.<br />

I


Summary II<br />

Overall, SPAR<strong>CM</strong>SG has proven to provide interesting new opportunities <strong>for</strong> <strong>cloud</strong><br />

detection using the MSG SEVIRI sensor. Results are particularly promising with<br />

regard to the application <strong>of</strong> SPARC <strong>for</strong> data <strong>of</strong> the MFG satellite series.


Contents<br />

Summary I<br />

Table <strong>of</strong> Contents III<br />

List <strong>of</strong> Figures V<br />

List <strong>of</strong> Tables VII<br />

1 Overview 1<br />

2 SPARC <strong>algorithm</strong> 3<br />

2.1 Cloud detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

2.2 Snow detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

2.3 Cloud shadow identification . . . . . . . . . . . . . . . . . . . . . . . 8<br />

3 The SPARC <strong>algorithm</strong> <strong>for</strong> MSG SEVIRI 9<br />

3.1 SPAR<strong>CM</strong>SG <strong>cloud</strong> mask . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

3.2 Calculation <strong>of</strong> the T-score . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

3.3 The SPAR<strong>CM</strong>SG snow detection scheme . . . . . . . . . . . . . . . . . 16<br />

3.4 Further modifications . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

4 Validation Study 22<br />

4.1 General Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

4.2 Data and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

III


Table <strong>of</strong> Contents IV<br />

4.2.1 ASRB data . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

4.2.2 MSG data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

4.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

4.3.1 Comparison with ASRB Cloud Parameters . . . . . . . . . . . 25<br />

4.3.2 MFG Simulation . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

4.3.3 Validation at Full Disk Level . . . . . . . . . . . . . . . . . . . 39<br />

4.4 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . 42


List <strong>of</strong> Figures<br />

2.1 SPARC output <strong>for</strong> a sample slot. . . . . . . . . . . . . . . . . . . . . 5<br />

3.1 HRV channel <strong>of</strong> a sample slot together with the SPARC output cre-<br />

ated with different thresholds. . . . . . . . . . . . . . . . . . . . . . . 11<br />

3.2 HRV channel <strong>of</strong> a sample slot together with the SPAR<strong>CM</strong>SG output. . 12<br />

3.3 Examples <strong>of</strong> a Mannstein function fitted to clear-sky BT108 values. . 15<br />

3.4 Example <strong>of</strong> a RST map in summer (left) and winter (right) over Swiss<br />

Alps. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

3.5 Output <strong>of</strong> the novel SPARC snow detection <strong>algorithm</strong>. . . . . . . . . 19<br />

3.6 Time series <strong>of</strong> monthly maximum snow cover in the Swiss Alps. . . . 20<br />

4.1 CFI vs. F <strong>for</strong> all sites and 12 months combined. . . . . . . . . . . . . 26<br />

4.2 CFI vs. F <strong>for</strong> each season separately. . . . . . . . . . . . . . . . . . . 26<br />

4.3 PCA vs. SPAR<strong>CM</strong>SG <strong>cloud</strong> mask <strong>for</strong> the entire 12 month period. . . . 28<br />

4.4 PCA vs. <strong>CM</strong>MSG <strong>for</strong> each season separately. . . . . . . . . . . . . . . 29<br />

4.5 Histogram <strong>of</strong> <strong>CM</strong>MSG output <strong>for</strong> each PCA class at the ASRB site in<br />

Payerne. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

4.6 Histogram <strong>of</strong> <strong>CM</strong>MSG output <strong>for</strong> each PCA class at the ASRB site at<br />

Weissflujoch. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

4.7 CFI vs. F <strong>for</strong> all sites and 12 months combined - SPAR<strong>CM</strong>FG mode. . 34<br />

4.8 Output <strong>of</strong> SPAR<strong>CM</strong>SG and SPAR<strong>CM</strong>FG on March 20, 2005 (8:45 am). 35<br />

4.9 Histogram <strong>of</strong> <strong>CM</strong>MFG output <strong>for</strong> each PCA class at the ASRB site in<br />

Payerne - SPAR<strong>CM</strong>FG mode. . . . . . . . . . . . . . . . . . . . . . . . 37<br />

V


List <strong>of</strong> Figures VI<br />

4.10 Histogram <strong>of</strong> <strong>CM</strong>MFG output <strong>for</strong> each PCA class at the ASRB site at<br />

Weissflujoch - SPAR<strong>CM</strong>FG mode. . . . . . . . . . . . . . . . . . . . . 38<br />

4.11 Comparison <strong>of</strong> the <strong>CM</strong>-<strong>SAF</strong>, SPAR<strong>CM</strong>SG, and SPAR<strong>CM</strong>FG <strong>cloud</strong> masks<br />

<strong>for</strong> the full disk. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

4.12 Comparison <strong>of</strong> the <strong>CM</strong>-<strong>SAF</strong>, SPAR<strong>CM</strong>SG, and SPAR<strong>CM</strong>FG <strong>cloud</strong> masks<br />

<strong>for</strong> five regions <strong>of</strong> interest within the full disk. . . . . . . . . . . . . . 41<br />

4.13 Modification <strong>of</strong> the R-score <strong>for</strong> bright desert surfaces. . . . . . . . . . 42


List <strong>of</strong> Tables<br />

2.1 Boundary coordinates <strong>of</strong> the region <strong>of</strong> interest. . . . . . . . . . . . . . 7<br />

4.1 Sites <strong>of</strong> the ASRB network used <strong>for</strong> the validation. . . . . . . . . . . 23<br />

4.2 Channels considered as SPARC input in MSG and MFG mode. . . . 25<br />

VII


List <strong>of</strong> Tables VIII


Chapter 1<br />

Overview<br />

It is <strong>of</strong> major importance <strong>for</strong> any satellite data processing system to accurately de-<br />

termine the state <strong>of</strong> a pixel. Depending on the application it may be necessary to<br />

assign a pixel to the clear-sky (e.g., <strong>for</strong> land surface applications) or <strong>cloud</strong>y cate-<br />

gory (e.g., <strong>for</strong> the retrieval <strong>of</strong> <strong>cloud</strong> properties). Further, assigning <strong>for</strong> instance a<br />

mixed pixel with sub-pixel <strong>cloud</strong> cover to the clear-sky or <strong>cloud</strong>y category largely<br />

depends on the final application. For land surface applications a clear sky con-<br />

servative approach is preferred, where <strong>for</strong> the retrieval <strong>of</strong> <strong>cloud</strong> properties a <strong>cloud</strong><br />

conservative approach has to be used. To achieve discrimination <strong>of</strong> various scene<br />

types (e.g., snow, <strong>cloud</strong>s, water, or vegetation), pixel scene identification <strong>algorithm</strong>s<br />

make use <strong>of</strong> the basic underlying principle that scene types may be distinguished<br />

based on their inherent reflective and emissive properties in certain portions <strong>of</strong> the<br />

electromagnetic spectrum (Lillesand et al., 2004).<br />

Scene identification <strong>algorithm</strong>s traditionally follow a branching/thresholding ap-<br />

proach: a sequence <strong>of</strong> spectral tests is applied to each pixel, assigning the pixel<br />

to a certain class (e.g., <strong>cloud</strong>y, clear-sky, or partly <strong>cloud</strong>y) based on predefined<br />

thresholds. However, such branching/thresholding approaches have two main dis-<br />

advantages (Khlopenkov and Trishchenko, 2007): first, since a yes/no decision is<br />

made at each node <strong>of</strong> the classification tree, a pixel that happens to represent an<br />

intermediate state may be assigned to the wrong category, or its classification may<br />

not be possible. Second, the final result <strong>of</strong> such classification schemes only contains<br />

in<strong>for</strong>mation on a limited number <strong>of</strong> classes, such as ’<strong>cloud</strong>y’, ’clear-sky’, ’partly<br />

<strong>cloud</strong>y’, or ’uncertain’, but does not provide in<strong>for</strong>mation on the degree <strong>of</strong> <strong>cloud</strong> or<br />

1


SPARC <strong>algorithm</strong> 2<br />

haze contamination within the field <strong>of</strong> view. Such in<strong>for</strong>mation may, however, be<br />

valuable <strong>for</strong> the generation <strong>of</strong> clear-sky composites <strong>for</strong> many applications, e.g., land<br />

surface albedo derivation or vegetation dynamics analysis.<br />

In order to address these limitations, the Separation <strong>of</strong> Pixels Using Aggregated<br />

Rating over Canada (SPARC) <strong>algorithm</strong> (Khlopenkov and Trishchenko, 2007) was<br />

developed at the Canada Centre <strong>for</strong> Remote Sensing (CCRS). The SPARC <strong>algorithm</strong><br />

was originally designed <strong>for</strong> the use with Advanced Very High Resolution Radiometer<br />

(AVHRR) data over Canada, but its design should theoretically enable the <strong>algorithm</strong><br />

to be transferred to other multi-spectral sensors such as the Meteosat Second Gen-<br />

eration (MSG) Spinning Enhanced Visible and Infrared Imager (SEVIRI) and other<br />

geographical regions <strong>of</strong> interest, even though some modification might be necessary.<br />

The aim <strong>of</strong> this report is to provide a short overview <strong>of</strong> the SPARC <strong>algorithm</strong> (Sec-<br />

tion 2) and the modifications specific <strong>for</strong> its application with the MSG SEVIRI sensor<br />

(Section 3). The report also presents the results <strong>of</strong> a validation study per<strong>for</strong>med<br />

over the Swiss Alps (Section 4.3.1). The applicability <strong>of</strong> the SPARC <strong>algorithm</strong> to<br />

data from the Meteosat First Generation (MFG) satellite series is evaluated (Section<br />

4.3.2). Finally, preliminary results <strong>of</strong> the applicaiton <strong>of</strong> SPARC at SEVIRI full disk<br />

dimension are presented (Section 4.3.3) and concluding remarks are given in Section<br />

4.4.


Chapter 2<br />

SPARC <strong>algorithm</strong><br />

2.1 Cloud detection<br />

In<strong>for</strong>mation provided in this chapter closely follows the work <strong>of</strong> Khlopenkov and<br />

Trishchenko (2007). The SPARC <strong>algorithm</strong> outputs a score, which can be regarded<br />

as a measure <strong>of</strong> the <strong>cloud</strong> contamination probability <strong>for</strong> each pixel. The SPARC<br />

score (F) is obtained through the summation <strong>of</strong> the output <strong>of</strong> a number <strong>of</strong> individual<br />

tests, which make use <strong>of</strong> all five spectral channels <strong>of</strong> the AVHRR sensor (Eq. 2.1):<br />

where<br />

F = T + B + R + �<br />

Ai, (2.1)<br />

• T is the temperature score, which is based on the comparison <strong>of</strong> the brightness<br />

temperature at 10.8µm and a dynamic threshold derived from an external sur-<br />

face skin temperature field. In the original SPARC version, surface skin tem-<br />

perature is obtained from the North American Regional Reanalysis (NARR)<br />

at a spatial resolution <strong>of</strong> 32 km×32 km (Mesinger and Coauthors, 2006).<br />

• B is the brightness score, calculated from the reflectance in the visible (VIS)<br />

and near infrared (NIR) portion <strong>of</strong> the electromagnetic spectrum over land<br />

and water, respectively,<br />

• R is the reflectance score calculated either from the reflectance values in the<br />

VIS and short wave infrared (SWIR) portions <strong>of</strong> the spectrum (if a SWIR<br />

3<br />

i


SPARC <strong>algorithm</strong> 4<br />

channel is available from AVHRR), or from the solar reflective component at<br />

3.7µm otherwise, and<br />

• Ai represents the added scores <strong>of</strong> four additional tests that are only per<strong>for</strong>med<br />

under certain circumstances due to computational considerations.<br />

These supplementary tests include:<br />

• the simple ratio test (N), exploiting the relatively homogeneous reflective prop-<br />

erties <strong>of</strong> <strong>cloud</strong>y scenes in the VIS and NIR spectrum as opposed to clear-sky<br />

observation <strong>of</strong>, e.g., vegetation cover,<br />

• the texture uni<strong>for</strong>mity test (Utext) based on the NIR spectrum and the ther-<br />

mal uni<strong>for</strong>mity test (Utemp) based on 10.8µm brightness temperature, which<br />

both analyze the variability <strong>of</strong> the pixel values <strong>for</strong> a pixel and its four nearest<br />

neighbors, and finally<br />

• the thin cirrus test based on the brightness temperature difference between<br />

12µm and 10.8µm.<br />

The scores are designed such that strong (weak) <strong>cloud</strong> evidence leads to positive<br />

(negative) values <strong>of</strong> F, whereas values <strong>of</strong> F ≈0 represent intermediate, i.e. uncertain<br />

states. In combination with the linear aggregation principle, this design has the<br />

advantage that a <strong>cloud</strong> mask can be generated even if some <strong>of</strong> the scores are not<br />

calculated. More detailed in<strong>for</strong>mation on the calculation <strong>of</strong> the scores is provided in<br />

Khlopenkov and Trishchenko (2007).<br />

However, some adjustments <strong>of</strong> Equation 2.1 are required in order <strong>for</strong> the single<br />

scores to receive appropriate weights under special observation conditions, namely<br />

snow cover, sun glint, and in the proximity to the terminator line. For this reason,<br />

three correction factors are implemented. These are:<br />

• the correction factor (s) <strong>for</strong> snow conditions, which is necessary since snow-<br />

covered areas observed under clear-sky conditions and <strong>cloud</strong>s can be equally<br />

bright. In the original SPARC scheme, this factor is estimated from the surface<br />

skin temperature fields;


SPARC <strong>algorithm</strong> 5<br />

• the sun glint correction factor (g; calculated from the observation geometry)<br />

to account <strong>for</strong> increased spectral reflectance over sun glint areas;<br />

• the nighttime correction factor (n), which accounts <strong>for</strong> the special circum-<br />

stances at very low Sun zenith angles in the proximity <strong>of</strong> the terminator line.<br />

Equation 2.1 is thus modified as follows:<br />

F = B(1 − g)sn + T (1 + g)(2 − s)(2 − n) + R(1 − 0.6g)(2 − s)n + N(1 − g)sn<br />

+ Utext(1 − g)sn + Utemp(1 + g)(2 − s)(2 − n) + C (2.2)<br />

Note that all correction factors reduce the weight <strong>of</strong> the reflective scores (where<br />

applicable <strong>for</strong> a certain correction factor) and assign more weight to the thermal<br />

tests. By design, reflectance scores are turned <strong>of</strong>f completely at nighttime and the<br />

SPARC output is based only on the test <strong>for</strong> the thermal channels. Figure 2.1 displays<br />

an example <strong>of</strong> all SPARC scores as well as the snow correction factor <strong>for</strong> a MSG<br />

SEVIRI time slot in March 2004. The boundary coordinates <strong>of</strong> the region <strong>of</strong> interest<br />

covering the Swiss Alps are provided in Table 2.1. The HRV channel (1A) shows<br />

significant snow cover over the elevated topography <strong>of</strong> the Swiss Alps under mostly<br />

clear-sky conditions as well as prevalent low stratus <strong>cloud</strong>s over the Swiss Main<br />

Plateau. Note that the absolute values <strong>of</strong> the single contributing scores vary greatly<br />

in magnitude. In particular, scores <strong>of</strong> the supplementary tests display significantly<br />

smaller absolute values compared to the three main scores (T, B, and R). Figure<br />

2.1 also highlights a problematic issue around the calculation <strong>of</strong> the Utext-scores<br />

over the Swiss Alps during wintertime: even under <strong>cloud</strong>-free conditions the high<br />

reflectance contrast between snow-covered mountain tops and low-lying and snow<br />

free valley bottoms results in increased values <strong>of</strong> the Utext-score. The same problem,<br />

Figure 2.1 (following page): High resolution visible (HRV) channel <strong>of</strong> the MSG SE-<br />

VIRI time slot on March 20, 2005 (8:45 am; 1A), together with the corresponding SPARC<br />

scores (see text <strong>for</strong> more details). The correction factor <strong>for</strong> snow conditions (s) is also<br />

provided (1E). Note that s was calculated as described in Khlopenkov and Trishchenko<br />

(2007), however, using a clear-sky BT108 mosaic as input (cp. Section 3.2) instead <strong>of</strong><br />

an external surface skin temperature map. Details on the calculation <strong>of</strong> the scores can be<br />

found in Khlopenkov and Trishchenko (2007).


SPARC <strong>algorithm</strong> 6


SPARC <strong>algorithm</strong> 7<br />

Table 2.1: Boundary coordinates <strong>of</strong> the region <strong>of</strong> interest (ROI) covering the Swiss Alps,<br />

where ’lat’ is the latitude [ ◦ ] and ’lon’ the longitude [ ◦ ].<br />

ROI Swiss Alps<br />

latmin<br />

latmax<br />

lonmin<br />

lonmax<br />

yet less pronounced, is observed <strong>for</strong> the Utemp-score. Similarly, the N-score adopts<br />

high values both <strong>for</strong> <strong>cloud</strong> and snow cover. Even though the contributions <strong>of</strong> these<br />

additional scores to the final score F are small, it is suspected that erroneously large<br />

values <strong>of</strong> these scores may lead to some misclassifications in uncertain conditions.<br />

2.2 Snow detection<br />

Snow detection is originally per<strong>for</strong>med in two steps based on the T-, B-, and R-<br />

scores:<br />

45.5<br />

48.0<br />

6.5<br />

10.0<br />

• In step one, a number <strong>of</strong> threshold tests is per<strong>for</strong>med, testing whether the pixel<br />

is sufficiently bright (B-score threshold), limiting the R-score to small values to<br />

exclude <strong>cloud</strong> contamination, and ensuring that the brightness temperature at<br />

10.8µm is close to the surface skin temperature map as outlined above. Once<br />

a pixel is classified as snow-covered in this first step, it undergoes another test<br />

in<br />

• step two, which is designed to detect <strong>cloud</strong>s over snow based on the R- and<br />

T-score. A pixel is classified as clear-sky snow, if both the R- and T-score are<br />

smaller than zero. The pixel is classified as snow-covered and <strong>cloud</strong> contami-<br />

nated, if these criteria are not met.<br />

Additional details as well as a flowchart <strong>of</strong> the SPARC snow <strong>algorithm</strong> can be found<br />

in Khlopenkov and Trishchenko (2007, Figures 4 and 5).


SPARC <strong>algorithm</strong> 8<br />

2.3 Cloud shadow identification<br />

Cloud shadow detection is per<strong>for</strong>med based on geometrical considerations, i.e. by<br />

computing the extent <strong>of</strong> the <strong>cloud</strong> shadow on the Earth’s surface (shadows cast<br />

by high <strong>cloud</strong>s on the top <strong>of</strong> low <strong>cloud</strong>s are not considered). The output mask<br />

includes in<strong>for</strong>mation on three different categories: shadow-free and, according to<br />

two thresholds to F, thin <strong>cloud</strong> shadow and thick <strong>cloud</strong> shadow. For more details<br />

on the <strong>cloud</strong> shadow retrieval it is again referred to Khlopenkov and Trishchenko<br />

(2007).


Chapter 3<br />

The SPARC <strong>algorithm</strong> <strong>for</strong> MSG<br />

SEVIRI<br />

Prior to the application <strong>of</strong> the SPARC <strong>algorithm</strong> to data from the MSG SEVIRI<br />

sensor (hereafter referred to as SPAR<strong>CM</strong>SG), a number <strong>of</strong> modifications relative to<br />

the original version (SPARCORIG) were conducted. These will be explained in the<br />

following sections.<br />

3.1 SPAR<strong>CM</strong>SG <strong>cloud</strong> mask<br />

Other than in SPARCORIG, the primary output <strong>of</strong> the SPAR<strong>CM</strong>SG <strong>cloud</strong> detection<br />

scheme is not the score F itself, but rather a <strong>cloud</strong> mask with three different classes:<br />

clear (0), thin <strong>cloud</strong> (1), and thick <strong>cloud</strong> (2). Similar to the criteria in Khlopenkov<br />

and Trishchenko (2007) <strong>for</strong> the distinction between shadows caused by optically<br />

thin or thick <strong>cloud</strong>s, respectively, this simplified <strong>cloud</strong> mask is obtained by applying<br />

thresholds to F:<br />

0 : F < ct : clear-sky; (3.1)<br />

1 : 4 > F > ct : thin <strong>cloud</strong>;<br />

2 : F > 4 : thick <strong>cloud</strong>,<br />

where ct is the <strong>cloud</strong> threshold, which <strong>for</strong> each pixel is calculated incrementally as<br />

ct = −4.0 : as a basic <strong>cloud</strong> threshold (3.2)<br />

9


SPARC <strong>algorithm</strong> 10<br />

ct = ct − 5.0s : stricter threshold if T-score was not applied<br />

ct = ct − 5.0s : even stricter if C-score was not applied.<br />

Even though SPARC was originally designed to overcome limitations <strong>of</strong> traditional<br />

<strong>cloud</strong> detection schemes as outlined above, the <strong>cloud</strong> mask generation based on the<br />

application <strong>of</strong> thresholds to F represents an interesting feature <strong>of</strong> the new SPAR<strong>CM</strong>SG<br />

<strong>algorithm</strong>: by changing the thresholds in either direction, the resulting <strong>cloud</strong> mask<br />

may be shifted dynamically from a clear-sky conservative to a <strong>cloud</strong> conservative<br />

mode without modifying the retrieval <strong>algorithm</strong> itself. This is illustrated in Figure<br />

3.1 which displays the high resolution visible (HRV) channel <strong>of</strong> a sample slot (A)<br />

covering the Swiss Alps together with F (B) and a number <strong>of</strong> <strong>cloud</strong> masks (D-<br />

I). Cloud masks were generated based on basic <strong>cloud</strong> thresholds varying between<br />

−8≤ct≤4 (thresholds <strong>for</strong> thick <strong>cloud</strong> were changed accordingly and were set to<br />

ct+8). The <strong>cloud</strong> mask displayed in Figure 3.1, panel F, was generated following<br />

Equations 3.1 and 3.2. Note that significant differences are observed depending on<br />

the selected threshold, especially in areas where the <strong>cloud</strong> mask error (see below)is<br />

large.<br />

In a subsequent step, the error (ε) <strong>of</strong> the <strong>cloud</strong> mask is calculated based on F as<br />

follows:<br />

(F −ε)2<br />

−<br />

ε = e (2σ2 ε ) , (3.3)<br />

where ε = 0.0<br />

σε = 10.0<br />

(F − ε) 2 � 2000.0<br />

The <strong>cloud</strong> mask error may adopt values between 0 and 1 is designed to be largest at<br />

values <strong>of</strong> F=0, i.e., in situations where the summation <strong>of</strong> the SPARC scores did not<br />

give clear evidence <strong>for</strong> either <strong>cloud</strong>y or clear-sky conditions. Hence, ε is a potentially<br />

useful parameter, e.g., to set up decision criteria in the clear-sky compositing process.<br />

Figure 3.2 displays the high resolution visible (HRV) channel <strong>of</strong> a MSG SEVIRI time<br />

slot in February, together with the corresponding SPARC score F, <strong>cloud</strong> mask, and<br />

<strong>cloud</strong> mask error ε. Cloud mask error is also displayed in Figure 3.1 (C).


SPARC <strong>algorithm</strong> 11<br />

Figure 3.1: The HRV channel <strong>of</strong> a sample slot on 11 September 2004, 11 am (A), together<br />

with the SPARC score (B), the <strong>cloud</strong> mask error (C; see text <strong>for</strong> explanations) and a<br />

number <strong>of</strong> <strong>cloud</strong> masks based on varying thresholds (C-H; see text <strong>for</strong> further explanations).<br />

Black areas represent clear-sky conditions, optically thick (thin) <strong>cloud</strong>s are delineated in<br />

white (grey).


SPARC <strong>algorithm</strong> 12<br />

Figure 3.2: The HRV channel <strong>of</strong> a sample slot channel on 25 February 2005, 10:15 am<br />

(top, left) together with the SPARC score (top, right), <strong>cloud</strong> mask (bottom, left), and <strong>cloud</strong><br />

mask error (bottom, right). Please note the increased values <strong>of</strong> the error ε along the <strong>cloud</strong><br />

boundaries.


SPARC <strong>algorithm</strong> 13<br />

3.2 Calculation <strong>of</strong> the T-score<br />

One <strong>of</strong> the major modifications with regard to SPARCORIG concerns the calculation<br />

<strong>of</strong> the T-score. The high temporal resolution <strong>of</strong> the SEVIRI sensor (15 minutes)<br />

opens up new opportunities <strong>for</strong> surface temperature retrievals, because the chance<br />

<strong>of</strong> a certain location being observed under clear-sky conditions during the course <strong>of</strong> a<br />

single day is significantly increased compared to the AVHRR sensor (1-2 observations<br />

daily). For this reason, an external surface temperature map is not implemented in<br />

SPAR<strong>CM</strong>SG. Instead, a clear sky radiative surface temperature (RST) map directly<br />

derived from SEVIRI 10.8µm channel radiances (BT108) is used as a reference in<br />

the calculation <strong>of</strong> the T-score. As a preliminary solution, brightness temperature is<br />

used instead <strong>of</strong> the radiative surface temperature, because atmospheric correction<br />

is not yet implemented.<br />

The RST map is generated daily as follows: after the processing <strong>of</strong> all 96 daily slots<br />

is completed, the BT108 clear-sky values <strong>for</strong> each pixel and slot are aggregated in<br />

a running array <strong>of</strong> clear sky BT108 values. In this running array, which stores a<br />

maximum <strong>of</strong> two clear-sky values per pixel and slot, the oldest clear-sky BT108<br />

values are replaced by the current observations. However, values observed more<br />

than ten days ago are removed even if there are no current clear-sky observations<br />

available <strong>for</strong> replacement. The resulting array is subsequently used to model the<br />

diurnal evolution <strong>of</strong> surface brightness temperature <strong>for</strong> each pixel on the following<br />

day. This is per<strong>for</strong>med by fitting a modified version <strong>of</strong> the diurnal surface radiative<br />

temperature model (f) by Mannstein et al. (1999). The model is defined as follows<br />

(Dürr, 2009):<br />

2<br />

−2(ωt − a3)<br />

f = a0 + a1(exp( ) + 0.1sin(ωt − a3)) (3.4)<br />

a 2 2<br />

where ω = 2π/96, with 96 the total number <strong>of</strong> slots per day, a0 is the minimium<br />

brightness temperature (BT) <strong>of</strong> the diurnal cycle in Kelvin, a1 is the diurnal ampli-<br />

tude <strong>of</strong> BT in Kelvin, a2 is the half width <strong>of</strong> the day lenght in radians, and a3 is the<br />

true Sun time in radians. Please consult Dürr (2009) <strong>for</strong> more detailed in<strong>for</strong>mation<br />

on the model. Points are weighted as follows, considering the <strong>cloud</strong> mask error <strong>for</strong><br />

the corresponding BT108 clear-sky value as well as the date <strong>of</strong> observation:<br />

w =<br />

1<br />

ε + t , (3.5)<br />

tmax


SPARC <strong>algorithm</strong> 14<br />

where w =weight <strong>for</strong> each point, t =the number <strong>of</strong> days the clear-sky BT108 value<br />

had been stored in the running mosaic, and tmax=10, i.e. the maximum number <strong>of</strong><br />

days considered in the clear-sky compositing (may be adjusted in the configurations<br />

file). Hence, smallest weights are assigned to BT108 values observed early in the<br />

preceeding compositing interval <strong>of</strong> tmax days and under uncertain conditions (F≈0).<br />

Figure 3.3 shows examples <strong>of</strong> a RST fit <strong>for</strong> a location on the Swiss Main Plateau<br />

in summer (left) and winter (right). Please note the difference in the evolution and<br />

the magnitude <strong>of</strong> diurnal brightness temperature.<br />

In addition, a spatial interpolation procedure is implemented in order to fill gaps<br />

in the RST field. The interpolation is based on an elevation-dependent RST filling<br />

using a lapse-rate that is retrieved from neighboring valid pixels with a radius <strong>of</strong> 50<br />

SEVIRI HRV pixels. Figure 3.4 provides an example <strong>of</strong> a RST map <strong>of</strong> the Swiss<br />

Alps in summer 2005 (left) and winter 2004/2005 (right).<br />

The use <strong>of</strong> a RST map derived directly from MSG data has two major advantages:<br />

first, the suggested SPAR<strong>CM</strong>SG <strong>cloud</strong> detection scheme represents a stand-alone ap-<br />

proach that does not rely on any auxillary input data, such as the coarse spatial<br />

resolution NARR surface skin temperature fields as described in Section 2. Sec-<br />

ond, in contrast to these modelled surface temperature fields, the RST map ap-<br />

parently provides substantial spatial detail, which is assumed to be beneficial <strong>for</strong><br />

<strong>cloud</strong> detection over the complex and heterogeneous terrain <strong>of</strong> the Swiss Alps. The<br />

disadvantage <strong>of</strong> modelling the following day’s diurnal BT108 amplitude from the<br />

previous day’s clear-sky BT108 observations is that day-to-day changes in BT108<br />

at a certain time <strong>of</strong> the day cannot be captured. Resulting biases in the T-score<br />

could translate into erroneous <strong>cloud</strong> masks, especially if such day-to-day changes<br />

happen to be significant. However, this problem is likely to occur only in the case <strong>of</strong><br />

warm and low <strong>cloud</strong>s, when the temperature difference between ground and <strong>cloud</strong><br />

top are small. Dectection <strong>of</strong> high and cold <strong>cloud</strong>s should not be affected. A solution<br />

might, however, be to apply a two-stage processing, where first the RST <strong>of</strong> the full<br />

day is calculated, then the <strong>cloud</strong> mask <strong>for</strong> the full day is repeated with the cur-<br />

rent day’s RST. Furthermore, this stand-alone approach implies that, under some<br />

circumstances, the T-score cannot be calculated <strong>for</strong> certain pixels. This is the case<br />

either at the beginning <strong>of</strong> the processing sequence, when the RST map is not defined<br />

due to the lack <strong>of</strong> clear-sky BT values in the mosaic, or if no clear-sky BT values


SPARC <strong>algorithm</strong> 15<br />

Figure 3.3: Examples <strong>of</strong> a Mannstein function fitted to clear-sky BT108 values <strong>for</strong> a<br />

location in the Swiss main land in summer (left) and winter (right). The error bars are<br />

calculated as defined in Equation 3.5.<br />

Figure 3.4: Example <strong>of</strong> a RST map in summer (left) and winter (right) over the Swiss<br />

Alps Region <strong>of</strong> Interest (ROI; see Table 2.1 <strong>for</strong> details on the ROI). Please note the bright-<br />

ness temperature contrast between the low-lying valleys and the more elevated topography,<br />

especially in winter.


SPARC <strong>algorithm</strong> 16<br />

had been observed <strong>for</strong> more than tmax days (certain gaps cannot be filled using the<br />

spatial interpolation scheme if large areas are affected). Limitations at the beginning<br />

<strong>of</strong> a processing sequence can be avoided by adding a spin-up time prior to the actual<br />

period <strong>of</strong> interest. Theoretically, one single clear-sky day would be sufficient to gen-<br />

erate a complete RST field. However, escpecially during the winter months when<br />

low stratus <strong>cloud</strong> cover over the Swiss Main Plateau is observed frequently, gaps in<br />

the RST field may still be observed after several weeks <strong>of</strong> observation and longer<br />

spin-up times are required. However, due to the design <strong>of</strong> the scores and the linear<br />

aggregation principle as mentioned above, SPAR<strong>CM</strong>SG can also be employed if not<br />

all scores are calculated. See Section 4.3.2 <strong>for</strong> a discussion <strong>of</strong> a situation where not<br />

all scores are calculated. Another option could be to consider high resolution surface<br />

temperature data from, e.g., the COSMO-2 model over the Swiss Alps (grid spacing<br />

<strong>of</strong> about 2.2 km; Doms and Schaettler, 2002) as a reference in T-score calculation<br />

where the RST map is not defined.<br />

3.3 The SPAR<strong>CM</strong>SG snow detection scheme<br />

The third major modification <strong>of</strong> the SPARC <strong>algorithm</strong> concerns the calculation <strong>of</strong><br />

the snow mask. A daily snow mask is generated through the aggregation <strong>of</strong> all<br />

snow-covered pixels observed under clear-sky conditions during the course <strong>of</strong> the<br />

day. Snow detection in SPAR<strong>CM</strong>SG includes a number <strong>of</strong> different tests, where<strong>of</strong><br />

one includes the calculation <strong>of</strong> the Normalized Difference Snow Index (NDSI). The<br />

NDSI ([−1.0,1.0]) is calculated by dividing the difference <strong>of</strong> reflectances observed<br />

in MSG SEVIRI channels 1 and 4 by their sum and may be regarded as a measure<br />

<strong>of</strong> the abundance <strong>of</strong> snow and ice within the area covered by a pixel (Salomonson<br />

and Appel, 2004). Even though the spectral channels needed <strong>for</strong> the calculation <strong>of</strong><br />

the NDSI are available on the AVHRR sensors <strong>of</strong> the third generation (AVHRR/3),<br />

historical AVHRR sensors lack the channel at 1.6µm. As a consequence, the NDSI<br />

cannot be implemented systematically <strong>for</strong> AVHRR, which is the reason why an<br />

alternative method was originally developed as described in Section 2.2.<br />

The idea behind the new snow detection scheme is the calculation <strong>of</strong> a S-score that<br />

can be regarded as a measure <strong>of</strong> the snow evidence within a pixel. The S-score is<br />

calculated by summing two newly defined scores, the NDSI-score and the FREEZE-


SPARC <strong>algorithm</strong> 17<br />

score, as well as the T- and R-scores as described above. The NDSI-score is defined<br />

as follows:<br />

NDSI =<br />

REF 006 − REF 016<br />

REF 006 + REF 016<br />

NDSI-score = (NDSI − 0.4)40.0, (3.6)<br />

where REF 006 is the reflectance in SEVIRI channel 1 at 0.6µm and REF 016 the<br />

reflectance in channel 3 at 1.6µm. Positive (negative) values <strong>of</strong> the NDSI-score are<br />

obtained <strong>for</strong> snow-covered (snow-free) pixels. The FREEZE-score is calculate as<br />

follows:<br />

FREEZE-score = −|BT 108 − (Tfreeze) + 5|, (3.7)<br />

where Tfreeze is calculated as a sine function oscillating between a maximmum <strong>of</strong><br />

+2 ◦ C in spring and a minimum <strong>of</strong> −2 ◦ C in fall (Khlopenkov and Trishchenko, 2007).<br />

The FREEZE-score is designed to adopt negative values if the difference between<br />

BT108 and Tfreeze is large (positive or negative), and positive values if BT108 is<br />

close to Tfreeze. The S-score is finally aggregated as<br />

S-score = (R − 3.0)(−1.0) + NDSI-score + (T − 3.0)(−1.0) + FREEZE-score (3.8)<br />

The snow mask is subsequently generated from the S-score by classifying pixels<br />

as snow-covered if the S-score is greater than zero. Again, the sensitivity <strong>of</strong> the<br />

<strong>algorithm</strong> may be adjusted by tuning the selected threshold. Snow masks are only<br />

generated <strong>for</strong> observations made at Sun zenith angles (SZAs) smaller than 75 ◦ , since<br />

the suggested approach in some cases failed to accurately detect snow cover in the<br />

proximity <strong>of</strong> the terminator line, mainly due to erroneously high avalues <strong>of</strong> the<br />

NDSI-score. In addition, two criteria are applied to the snow mask:<br />

• For pixels where BT>tfreeze+10.0, the snow mask is set to zero. This threshold<br />

is needed to account <strong>for</strong> subpixel snow in mountains in summer.<br />

• For pixels with reflectance values in the visible channel (HRV or R0.6) <strong>of</strong>


SPARC <strong>algorithm</strong> 18<br />

with Smin=−15.0 and Smax=10.0. Given the limitations <strong>of</strong> the new snow detection at<br />

high SZAs, the snow correction factor s must be adjusted <strong>for</strong> SZAs greater than 75 ◦ .<br />

The preliminary implementation <strong>of</strong> this correction at high SZAs reads as follows:<br />

tf = e SZA−75.0<br />

10.0<br />

s75 = (s + 0.3)tf, (3.10)<br />

where tf is the correction factor in the proximity <strong>of</strong> the terminator line and s75 is<br />

the adjusted snow correction factor. In order to overcome limitations due to low<br />

SZAs, a possible future solution could be not to generate snow masks at all <strong>for</strong><br />

observations made at SZA>75 ◦ , or, adopting the idea <strong>of</strong> the nighttime correction<br />

factor, to reduce (increase) the weights <strong>of</strong> the reflective (thermal) scores in Equation<br />

3.8 in the proximity to the terminator line.<br />

Figure 3.5 displays the single scores <strong>of</strong> the new snow detection scheme (1A, 2A,1B),<br />

the resulting snow mask (1C) as well as the corresponding snow correction factor<br />

s (2B) <strong>for</strong> a day in March 2005. See Figure 4.8 (left) <strong>for</strong> additional SPAR<strong>CM</strong>SG<br />

output <strong>for</strong> the same time slot. The corresponding snow mask generated based on<br />

the original snow detection scheme is provided as a reference (2C).<br />

Major differences are observed both <strong>for</strong> s and the daily snow mask. Given the<br />

importance <strong>of</strong> the snow correction factor s to reduce the weight <strong>of</strong> the reflective<br />

scores in Equation 2.2, such differences in s inevitably translate into discrepancies<br />

in the magnitude <strong>of</strong> F in SPAR<strong>CM</strong>SG compared to SPARCORIG. Comparing the<br />

values <strong>of</strong> F from SPARCORIG in Figure 2.1 (2E) with SPAR<strong>CM</strong>SG F in Figure 4.8<br />

(1E), significant differences become apparent in the Swiss Alps. Panel 2E in Figure<br />

2.1 shows increased values <strong>of</strong> F over snow-covered areas, except where reflective<br />

scores were turned <strong>of</strong> by s (seen as dark areas; see also Panel 1E in Figure 2.1).<br />

Regarding the comparison <strong>of</strong> the daily snow masks, from a purely visual inspection<br />

the novel snow mask seems to be more accurate, with snow cover being understi-<br />

mated by SPARCORIG. Figure 3.6 (left) displays a preliminary time series <strong>of</strong> monthly<br />

snow cover from SPAR<strong>CM</strong>SG, SPARCORIG, and Moderate Resolution Imaging Spec-<br />

troradiometer (MODIS) data over the Swiss Alps. Monthly maximum snow cover<br />

composites <strong>for</strong> the period between July 2004 and June 2005 were generated <strong>for</strong><br />

both SPARC versions by considering all pixels that were classified as ’snow’ at any<br />

time during the month. Monthly MODIS snow composites provide average monthly


SPARC <strong>algorithm</strong> 19<br />

Figure 3.5: Output <strong>of</strong> the novel SPARC snow detection <strong>algorithm</strong> on March 20, 2005<br />

(8:45am). The corresponding snow mask generated based on the original snow detection<br />

scheme is provided as a reference (2C). See Figure 4.8 <strong>for</strong> the corresponding HRV image<br />

and SPAR<strong>CM</strong>SG.


SPARC <strong>algorithm</strong> 20<br />

Figure 3.6: Time series <strong>of</strong> monthly maximum snow-covered area (SCA; [%]) in the Swiss<br />

Alps region <strong>of</strong> interest <strong>for</strong> SPARCORIG, SPAR<strong>CM</strong>SG, and MODIS (see the text <strong>for</strong> more<br />

in<strong>for</strong>mation on the generation <strong>of</strong> the composites).<br />

snow cover <strong>for</strong> a certain pixel. For comparison with SPARC SCA, MODIS data<br />

were trans<strong>for</strong>med into a binary snow mask by classifying a pixel as snow-covered<br />

if the average monthly snow cover was ≥50%. Differences [%] relative to MODIS<br />

are provided in the right panel. An examination <strong>of</strong> SCA time series reveals that<br />

the underestimation <strong>of</strong> SCA by SPARCORIG relative to SPAR<strong>CM</strong>SG is observed in-<br />

dependently <strong>of</strong> the season. Relative to MODIS, SPAR<strong>CM</strong>SG underestimates (overes-<br />

timates) SCA in summer (fall, winter, and spring), whereas SPARCORIG SCA does<br />

not follow the same pattern. During fall, winter, and spring, an overestimation <strong>of</strong><br />

SCA by SPAR<strong>CM</strong>SG is in accordance with theory since ephemeral snow cover may<br />

not be captured by MODIS due to the lower temporal resolution <strong>of</strong> the MODIS<br />

sensor system. However, during summer we would expect the discrepancies to be<br />

small. Further analysis is required to explain the underestimation <strong>of</strong> SCA by both<br />

SPARC <strong>algorithm</strong>s.<br />

3.4 Further modifications<br />

In the original version <strong>of</strong> SPARC the additional scores are only calculated in uncer-<br />

tain situations to obtain more in<strong>for</strong>mation on the state <strong>of</strong> the pixel. As <strong>of</strong> now, this


SPARC <strong>algorithm</strong> 21<br />

distinction is not made in SPAR<strong>CM</strong>SG and all scores are calculated <strong>for</strong> each pixel<br />

and fed into Equation 2.2.<br />

Furthermore, the coefficients <strong>for</strong> the calculation <strong>of</strong> the R-score needed to be modified,<br />

since the contour lines in Khlopenkov and Trishchenko (2007, Figure 3) could not be<br />

reproduced with MSG SEVIRI reflectances using the original coefficients. The new<br />

coefficients <strong>for</strong> the calculation <strong>of</strong> the R-score are: <strong>of</strong>fset=0.02 and scale factor=320<br />

(original coefficients: <strong>of</strong>fset=0.1, scale factor=160).


Chapter 4<br />

Validation Study<br />

4.1 General Remarks<br />

The focus <strong>of</strong> the validation study was set on a region <strong>of</strong> interest (ROI) covering the<br />

Swiss Alps. The boundary coordinates <strong>of</strong> the subset are listed in Table 2.1. A time<br />

period <strong>of</strong> 12 months (from July 1, 2004 to June 30, 2005) was considered. Further-<br />

more, seasons were analyzed independently, where winter included the months from<br />

December to February, spring from March to May, summer from June to August,<br />

and fall from September to November. In addition, differences between the Climate<br />

Monitoring (<strong>CM</strong>) Satellite Application Facility (<strong>SAF</strong>) <strong>cloud</strong> mask (EUMETSAT <strong>CM</strong><br />

<strong>SAF</strong> , 2009; Schulz et al., 2009) and SPAR<strong>CM</strong>SG were visually analyzed at full disk<br />

level <strong>for</strong> the 12 UTC slot on June 13, 2006, which is the day selected <strong>for</strong> the <strong>cloud</strong><br />

mask intercomparison study <strong>of</strong> the EUMETSAT workshop (Walther et al., 2009).<br />

4.2 Data and Methods<br />

4.2.1 ASRB data<br />

In<strong>for</strong>mation on <strong>cloud</strong> coverage derived from data <strong>of</strong> the Alpine Surface Radia-<br />

tion Budget (ASRB) network (Philipona et al., 1996; Marty and Philipona, 2000),<br />

was used <strong>for</strong> the validation <strong>of</strong> the SPAR<strong>CM</strong>SG <strong>cloud</strong> mask. Two parameters were<br />

compared to SPAR<strong>CM</strong>SG output at each ASRB site in the Swiss Alps (Dürr and<br />

22


SPARC <strong>algorithm</strong> 23<br />

Philipona, 2004): the Cloud-Free Index (CFI) derived from downward longwave ra-<br />

diation, temperature, and relative humidity in the Swiss Alps, and Partial Cloud<br />

Amount (PCA), which is derived from the CFI. The CFI adopts values >1.0 <strong>for</strong><br />

<strong>cloud</strong> cover and values


SPARC <strong>algorithm</strong> 24<br />

computational considerations, a shorter spin-up time <strong>of</strong> only 5 days was selected <strong>for</strong><br />

the full disk processing.<br />

The SPARC <strong>algorithm</strong> was run both in a MSG as well as a MFG mode. While<br />

the MSG mode includes five spectral channels, the MFG mode only relies on MSG<br />

channels 12 and 9 in order to simulate the spectral properties <strong>of</strong> the MFG sensor<br />

(see Table 4.2 <strong>for</strong> more in<strong>for</strong>mation). The restriction to only two spectral channels<br />

aimed at investigating the <strong>algorithm</strong>’s capability to detect <strong>cloud</strong>s with only a limited<br />

amount <strong>of</strong> spectral in<strong>for</strong>mation, analysis that may be useful with regard to the re-<br />

processing <strong>of</strong> MFG long-term data records. While differences to the MSG mode<br />

should be small at nighttime (when both <strong>algorithm</strong>s rely on thermal scores only),<br />

we would expect some discrepancies during daytime, when in<strong>for</strong>mation from all<br />

scores except the B-, UTEMP, and T-scores is not available in the MFG mode.<br />

As listed in Table 4.2, channels 12 (HRV) and 1 can be used interchangeably in MSG<br />

mode. If the HRV channel is selected, all other bands are resampled from native<br />

0.06 ◦ ×0.06 ◦ to the higher spatial resolution <strong>of</strong> the HRV channel (0.02 ◦ ×0.02 ◦ ). This<br />

implies that scores calculated using the HRV channel (Table 4.2) are influenced by<br />

eventual sub-grid <strong>cloud</strong>s that are not represented in the remaining scores to the same<br />

extent. For the analysis presented here the HRV channel was selected, neglecting<br />

uncertainties due to subgrid-scale features.<br />

For the comparison <strong>of</strong> ASRB measurements with SPARC output, a 9 pixels ×<br />

9 pixels area around each ASRB site was considered. At a spatial resolution <strong>of</strong><br />

0.02 ◦ ×0.02 ◦ this approximately corresponds to an area <strong>of</strong> 15×15 km 2 . As <strong>for</strong> the<br />

comparison with CFI, all values <strong>of</strong> F within the subset were averaged to obtain a<br />

single value <strong>of</strong> F per time slot and ASRB site. In order to compare the <strong>cloud</strong> mask<br />

with ASRB PCA, <strong>for</strong>mer was trans<strong>for</strong>med into partial <strong>cloud</strong> amount in ninths <strong>of</strong><br />

the 15×15 km 2 area. This was achieved by counting the number <strong>of</strong> pixels classified<br />

as ’<strong>cloud</strong>’ (thin and opaque) within the 9 pixels × 9 pixels subset. The resulting<br />

82 classes were then stratified into 10 classes <strong>of</strong> increasing <strong>cloud</strong> contamination (the<br />

0/9 class only contained entirely <strong>cloud</strong> free, i.e. 0/81, scenes).


SPARC <strong>algorithm</strong> 25<br />

Table 4.2: The spectral channels <strong>of</strong> SEVIRI used as input <strong>for</strong> SPARC in MSG and MFG<br />

mode, together with the SPARC scores associated with each spectral channel. Channels<br />

12 (HRV) and 1 can be used interchangeably in MSG mode, however, all channels are<br />

resampled to higher spatial resolution (0.02 ◦ ×0.02 ◦ ) if the HRV channel is selected instead<br />

<strong>of</strong> Channel 1.<br />

Mode<br />

Channel involved scores MSG MFG<br />

HRV (12) B (over land), N X X<br />

1 B (over land), N (X) -<br />

2 B (over water), N, Utext X -<br />

3 R X -<br />

9 T, Utemp, C X X<br />

10 C X -<br />

4.3 Results and Discussion<br />

4.3.1 Comparison with ASRB Cloud Parameters<br />

CFI vs. F<br />

Figure 4.1 displays a scatter-density plot <strong>for</strong> the comparison <strong>of</strong> CFI with F <strong>for</strong> day<br />

(left) and night observations (right; all ASRB sites and months were combined). A<br />

Sun zenith angle (SZA) threshold <strong>of</strong> 88 ◦ was used to separate day and night obser-<br />

vations. The dotted lines indicate the thresholds used <strong>for</strong> CFI (Section 4.2.1) and F<br />

(Section 3.1), respectively, to distinguish between <strong>cloud</strong>y and clear-sky observations<br />

(values <strong>of</strong> CFI


SPARC <strong>algorithm</strong> 26<br />

Figure 4.1: Scatter-density plot <strong>of</strong> CFI vs. F <strong>for</strong> all sites and 12 months combined <strong>for</strong><br />

day (left) and night observations (right). Dotted lines indicate the thresholds used <strong>for</strong><br />

CFI (Section 4.2.1) and F (Section 3.1), respectively, to distinguish between <strong>cloud</strong>y and<br />

clear-sky observations. Colors indicate the frequency <strong>of</strong> occurence <strong>for</strong> a certain CFI/F<br />

combination. A bin size <strong>of</strong> 0.07 (15.0) was selected <strong>for</strong> CFI (F); r: linear correlation<br />

coefficient.<br />

and night. Very large positive values <strong>of</strong> F are also observed, yet at very small<br />

frequencies. The agreement <strong>for</strong> nighttime observations was clearly weaker (r 2 =0.22).<br />

For values <strong>of</strong> CFI>1.0, F mostly adopts slightly negative values, which is indicative<br />

<strong>of</strong> an underestimation <strong>of</strong> nighttime <strong>cloud</strong> cover by SPAR<strong>CM</strong>SG, even though Dürr and<br />

Philipona (2004) report that the application <strong>of</strong> the CFI was found to overestimate<br />

nighttime <strong>cloud</strong> cover. A comparison with synoptic measurements will be needed to<br />

gain additional and more detailed insight into nighttime perormance <strong>of</strong> SPAR<strong>CM</strong>SG.<br />

Similar to Figure 4.1, the CFI/F relationship is shown in Figure 4.2, but <strong>for</strong> each<br />

season separately as defined in Section 4.1. For daytime observations, a strong<br />

Figure 4.2 (following page): Similar to Figure 4.1, but <strong>for</strong> each season separately.


SPARC <strong>algorithm</strong> 27


SPARC <strong>algorithm</strong> 28<br />

Figure 4.3: Observation frequencies <strong>of</strong> all possible PCA (in eights)/<strong>CM</strong>MSG (in ninths)<br />

combinations <strong>for</strong> the 12 month validation period during daytime (left) and nighttime<br />

(right). Colors indicate the frequency <strong>of</strong> occurence <strong>for</strong> a certain PCA/<strong>CM</strong>MSG combi-<br />

nation; corresponding percentage values are also provided.<br />

DFI/F relationship was observed in all seasons, even though better agreements were<br />

found <strong>for</strong> winter and spring (r 2 =0.59 and r 2 =0.61, respectively). Similar to Figure<br />

4.1, predominantly negative values <strong>of</strong> F are observed where CFI>1.0, showing that<br />

the underestimation <strong>of</strong> nighttime <strong>cloud</strong> cover by SPAR<strong>CM</strong>SG relative to ASRB CFI<br />

is observed independently <strong>of</strong> the season.<br />

PCA vs. SPAR<strong>CM</strong>SG Cloud Mask<br />

Two different analyses were per<strong>for</strong>med in order to validate the SPAR<strong>CM</strong>SG <strong>cloud</strong><br />

mask (hereafter referred to as <strong>CM</strong>MSG) with ASRB PCA. Firstly, frequencies <strong>of</strong><br />

occurence <strong>of</strong> certain PCA (in eights)/<strong>CM</strong>MSG (in ninths) <strong>cloud</strong> cover combinations<br />

were analyzed (Figures 4.3 and 4.4). Secondly, the accuracy <strong>of</strong> <strong>CM</strong>MSG was ana-<br />

lyzed separately <strong>for</strong> specific ASRB sites (Figure 4.5 and 4.6). Interestingly, Figure<br />

4.3 shows that in the majority <strong>of</strong> the cases during daytime (left) and nighttime<br />

(right), either complete clear-sky (PCA: 0/8, <strong>CM</strong>MSG: 0/9) or complete <strong>cloud</strong> cov-<br />

erage (PCA: 8/8, <strong>CM</strong>MSG: 9/9) was observed at the ASRB sites. Partial <strong>cloud</strong>


SPARC <strong>algorithm</strong> 29<br />

overcast is only observed in a limited number <strong>of</strong> cases, with


SPARC <strong>algorithm</strong> 30


SPARC <strong>algorithm</strong> 31<br />

plots suggest a clear-sky conservative behavior <strong>of</strong> SPAR<strong>CM</strong>SG, meaning that a pixel<br />

is only classified as ’clear-sky’ where <strong>cloud</strong> evidence is very low. In partial <strong>cloud</strong> cover<br />

situations, there is likely to be some <strong>cloud</strong> evidence at a certain pixel if the neigh-<br />

boring pixel is <strong>cloud</strong> covered, which eventually causes <strong>CM</strong>MSG to assign the pixel<br />

to the ”thin <strong>cloud</strong>” or ”opaque <strong>cloud</strong>” category. This is reflected in the tendency<br />

to classify partial <strong>cloud</strong> cover as completely <strong>cloud</strong> covered. At nighttime, results<br />

<strong>for</strong> partial <strong>cloud</strong> cover exhibit a more bimodal distribution, with partial <strong>cloud</strong> cover<br />

being classified as either clear-sky or complete overcast in most cases. Results <strong>for</strong><br />

the high elevation site at Weissfluhjoch (Figure 4.6) were similar, however, the per-<br />

centage <strong>of</strong> accurately classified daytime clear-sky scenes (70%) was lower compared<br />

to the Payerne site. Furthermore, clear-sky scenes were found to be misclassified as<br />

complete overcast in 8.4% <strong>of</strong> the cases during daytime.<br />

Figure 4.5 (following page): Histogram <strong>of</strong> <strong>CM</strong>MSG output <strong>for</strong> each PCA class <strong>for</strong><br />

daytime (top) and nighttime observations (bottom). Numbers represent relative frequencies<br />

[%]. Results are shown <strong>for</strong> the ASRB site in Payerne (cp. Table 4.1).


SPARC <strong>algorithm</strong> 32


SPARC <strong>algorithm</strong> 33<br />

Figure 4.6: Similar to Figure 4.5, but <strong>for</strong> the ASRB site at Weissfluhjoch (cp. Table<br />

4.1).


SPARC <strong>algorithm</strong> 34<br />

Figure 4.7: Similar to Figure 4.1 (Section 4.3.1), but <strong>for</strong> SPAR<strong>CM</strong>FG mode.<br />

4.3.2 MFG Simulation<br />

Similar to Figure 4.1 in Section 4.3.1, Figure 4.7 displays the comparison <strong>of</strong> CFI<br />

and F <strong>for</strong> the entire 12 month period, however, <strong>for</strong> SEVIRI data processed in the<br />

SPAR<strong>CM</strong>FG mode (see Section 4.2.2 <strong>for</strong> more details). As expected from theory,<br />

differences between both SPARC modes are minor during nighttime. Again, low<br />

values <strong>of</strong> F at values <strong>of</strong> CFI>1.0 are indicative <strong>of</strong> an underestimation <strong>of</strong> nighttime<br />

<strong>cloud</strong> cover. For daytime observations the relationship between CFI and F is weaker<br />

<strong>for</strong> SPAR<strong>CM</strong>FG, with r 2 =0.42 (r 2 =0.52 <strong>for</strong> SPAR<strong>CM</strong>SG; Figure 4.1).<br />

Note that daytime F adopts smaller positive values <strong>for</strong> <strong>cloud</strong>y situations in SPAR<strong>CM</strong>FG<br />

mode due to the smaller number <strong>of</strong> contributing scores. This becomes clear with a<br />

more detailed look at a SEVIRI sample time slot (8:45 am) on March 20, 2005 (Fig-<br />

ure 4.8). The RST mosaics <strong>of</strong> both SPARC modes (B1 and B2) generated based on<br />

clear-sky observations <strong>of</strong> the previous 10 days are quite similar and result in similar<br />

T-scores (C1 and C2; see also Section 2). However, since the R-score is not available<br />

in SPAR<strong>CM</strong>FG, the final SPARC scores (F) clearly differ in magnitide (E1 and E2),<br />

which translates into significantly different <strong>cloud</strong> masks (F1 and F2). Not only does


SPARC <strong>algorithm</strong> 35<br />

SPAR<strong>CM</strong>FG classify larger areas as clear-sky compared to SPAR<strong>CM</strong>SG, but it also<br />

assigns more <strong>cloud</strong>s to the ”thin <strong>cloud</strong>” category.<br />

Following the analysis in Section 4.3.1, the histograms <strong>of</strong> <strong>cloud</strong> mask output in<br />

SPAR<strong>CM</strong>FG mode (<strong>CM</strong>MFG) <strong>for</strong> each PCA class is displayed in Figures 4.9 and 4.10<br />

<strong>for</strong> the ASRB sites in Payerne and Weissfluhjoch, respectively. At the Payerne<br />

site, <strong>cloud</strong> cover appeared to be underestimated by SPAR<strong>CM</strong>FG not only during<br />

nighttime, but also during daytime, with 15% <strong>of</strong> the PCA 8/8 situations being<br />

classified as clear-sky. This is in contrast to the results in Figure 4.5, where a very<br />

good agreement between PCA (8/8) and <strong>CM</strong>MSG (9/9) was demonstrated (Section<br />

4.3.1). The same pattern, yet less pronounced, was observed at the Weissfluhjoch<br />

site, with a misclassification (PCA 8/8 vs. <strong>CM</strong>MFG 9/9) in 3.2% <strong>of</strong> the cases.<br />

Figure 4.8 (following page): Output <strong>of</strong> SPAR<strong>CM</strong>SG (B1-F1) and SPAR<strong>CM</strong>FG (B2-F2)<br />

on March 20, 2005 (8:45 am). The HRV channel (A1) and B-score (A2) are displayed<br />

once as they are identical <strong>for</strong> both SPARC configurations. R-score is not available in<br />

SPAR<strong>CM</strong>FG mode (D2), which results in significantly different <strong>cloud</strong> masks (0: clear-sky,<br />

1: thin <strong>cloud</strong>, 2: thick <strong>cloud</strong>). See the text <strong>for</strong> further explanations.


SPARC <strong>algorithm</strong> 36


SPARC <strong>algorithm</strong> 37<br />

Figure 4.9: Similar to Figure 4.5, but <strong>for</strong> SPAR<strong>CM</strong>FG mode.


SPARC <strong>algorithm</strong> 38<br />

Figure 4.10: Similar to Figure 4.6, but <strong>for</strong> SPAR<strong>CM</strong>FG mode.


SPARC <strong>algorithm</strong> 39<br />

4.3.3 Validation at Full Disk Level<br />

Similar to Figure 1 in Walther et al. (2009) we compared the output <strong>of</strong> SPAR<strong>CM</strong>SG<br />

and SPAR<strong>CM</strong>FG at full disk level to the <strong>cloud</strong> mask produced at the Climate Mon-<br />

itoring Satellite Application Facility (<strong>CM</strong>-<strong>SAF</strong>). The 12:00 UTC time slot on June<br />

13 th 2006 was selected. Prior to comparison, SPAR<strong>CM</strong>SG and SPAR<strong>CM</strong>FG output<br />

was trans<strong>for</strong>med into binary <strong>cloud</strong> masks by combining thin and opaque <strong>cloud</strong>s into<br />

one single class. Similarly, the <strong>CM</strong><strong>SAF</strong> <strong>cloud</strong> in<strong>for</strong>mation was converted to a binary<br />

<strong>cloud</strong> mask as described in Walther et al. (2009). The percentage <strong>of</strong> total <strong>cloud</strong> cover<br />

was then determined. The resulting <strong>cloud</strong> masks are displayed in Figure 4.11. From<br />

a visual comparison <strong>cloud</strong> cover looks quite similar <strong>for</strong> SPAR<strong>CM</strong>SG and SPAR<strong>CM</strong>FG,<br />

which is also confirmed by <strong>cloud</strong> cover percentage <strong>of</strong> 43% and 42%, respectively.<br />

The <strong>CM</strong>-<strong>SAF</strong> <strong>cloud</strong> mask estimates a <strong>cloud</strong> coverage <strong>of</strong> 55% <strong>of</strong> the full disk.<br />

In order to get more detailed insight into SPARC per<strong>for</strong>mance at full disk level,<br />

five subsets were extracted from the full disk, some <strong>of</strong> them similar to the ROIs<br />

in Walther et al. (2009): Central Europe, North Atlantic, Sahara Desert, Western<br />

Africa, and Southern Atlantic. Results are reported in Figure 4.12. At ROI level,<br />

larger differences between SPAR<strong>CM</strong>SG and SPARGMFG are visible. SPAR<strong>CM</strong>FG dis-<br />

plays a higher <strong>cloud</strong> cover percentage than SPAR<strong>CM</strong>SG, however, except <strong>for</strong> the<br />

Sahara ROI, both SPARC <strong>algorithm</strong>s detect significantly less <strong>cloud</strong> cover compared<br />

to the <strong>CM</strong>-<strong>SAF</strong> <strong>cloud</strong> mask. In the Sahara ROI, the extent <strong>of</strong> bright desert surfaces<br />

classified as ’<strong>cloud</strong>’ by both SPARC modes is significant. The reason <strong>for</strong> the over-<br />

estimation <strong>of</strong> <strong>cloud</strong> cover in these areas is tw<strong>of</strong>old: as <strong>for</strong> SPAR<strong>CM</strong>SG, it is mainly<br />

the very large values <strong>of</strong> the R-score over bright desert surfaces that translate into<br />

erroneous values <strong>of</strong> F, since the characteristics <strong>of</strong> the R-score do not account <strong>for</strong> the<br />

spectral properties <strong>of</strong> bright barren land. A possible approach could be to modify<br />

the R-score (Figure 4.13) or to introduce a weighting factor <strong>for</strong> barren land (similar<br />

to s <strong>for</strong> snow conditions). The <strong>for</strong>mer approach was implemented <strong>for</strong> the full disk<br />

processing presented here, however, Figure 4.12 suggests that further improvements<br />

are required. Latter approach may be considered <strong>for</strong> SPAR<strong>CM</strong>FG, as misclassifica-<br />

tion is mainly caused by strongly positive B-scores (R-score cannot be calculated in<br />

MFG mode). The positive B-scores over desert in MFG (and MSG) might also be


SPARC <strong>algorithm</strong> 40<br />

Figure 4.11: Comparison <strong>of</strong> the <strong>CM</strong>-<strong>SAF</strong> (top, right), SPAR<strong>CM</strong>SG (bottom, left), and<br />

SPAR<strong>CM</strong>FG (bottom, right) <strong>cloud</strong> masks <strong>for</strong> the full disk on 13 June, 2008, 12:00 UTC.<br />

The corresponding HRV channel is displayed in the top left panel. The <strong>cloud</strong> mask data<br />

was trans<strong>for</strong>med into a binary <strong>cloud</strong> mask (see the text <strong>for</strong> further explanations). The red<br />

squares (A-E) delineate the selected regions <strong>of</strong> interest (cp. Figure 4.12). Percent values<br />

indicate the <strong>cloud</strong> cover percentage on the full disk. Please note that the top and bottom<br />

212 lines were excluded from analysis due to computational constraints, which is also the<br />

reason why <strong>cloud</strong> cover percentage <strong>for</strong> the <strong>CM</strong>-<strong>SAF</strong> <strong>cloud</strong> mask differs from the value<br />

provided in Walther et al. (2009). Color legend: green: clear-sky land; blue: clear-sky<br />

water surface; yellow: <strong>cloud</strong>.


SPARC <strong>algorithm</strong> 41<br />

Figure 4.12: Comparison <strong>of</strong> the <strong>CM</strong>-<strong>SAF</strong> (second row), SPAR<strong>CM</strong>SG (third row), and<br />

SPAR<strong>CM</strong>FG (fourth row) <strong>cloud</strong> masks <strong>for</strong> five selected Regions <strong>of</strong> Interest (ROIs) within<br />

the full disk. Corresping HRV images are provided in the top row. The ROIs are: A<br />

- Central Europe, B - Northern Atlantic, C - Sahara Desert, D - Western Africa, E<br />

- Southern Atlantic (see also Figure 4.11). Percent values indicate the <strong>cloud</strong> coverage<br />

within the ROI.<br />

overcome by use <strong>of</strong> a similar score <strong>for</strong>mula as <strong>for</strong> the T-score:<br />

B-score = (V IS − ALB − <strong>of</strong>fset)scale (4.1)<br />

instead <strong>of</strong>: B-score = (V IS − <strong>of</strong>fset)scale,<br />

where ALB is the surface albedo and VIS the reflectance in Channel 1 (REF006)<br />

or 12 (HRV). This would mean that a spin-up <strong>for</strong> the surface albedo similar to the<br />

spin-up <strong>of</strong> the surface radiative temperature would need to be implemented. This<br />

would then make the snow-factor obsolete <strong>for</strong> the B-score. The surface albedo spin-<br />

up might, however, be hampered by rapidly changing surface albedo after snowfall<br />

periods.


SPARC <strong>algorithm</strong> 42<br />

Figure 4.13: Similar to Figure 3 in Khlopenkov and Trishchenko (2007), but only <strong>for</strong> the<br />

reflectance in SEVIRI channels 1 (REF006) and 3 (REF016) over bright desert surfaces.<br />

The red line demonstrates how the R-score should be modified in order to account <strong>for</strong> the<br />

spectral properties <strong>of</strong> bright desert surfaces.<br />

4.4 Concluding Remarks<br />

The Separation <strong>of</strong> Pixels Using Aggregated Rating Over Canada (SPARC) algo-<br />

rithm represents an alternative to conventional scene identification <strong>algorithm</strong>s as<br />

it primarily outputs a single <strong>cloud</strong> contamination rating (F) instead <strong>of</strong> a <strong>cloud</strong><br />

mask with a limited number <strong>of</strong> categories. The rating itself isessentially calculated<br />

from three sub-scores generated from in<strong>for</strong>mation on the brightness temperature<br />

(T-score), brightness (B-score), and the reflectance in the visible and short-wave<br />

infrared portion <strong>of</strong> the electromagnetic spectrum (R-score). Even though the algo-<br />

rithm was originally designed <strong>for</strong> AVHRR, it may be applied to data from other<br />

multispectral sensors.


SPARC <strong>algorithm</strong> 43<br />

The application <strong>of</strong> SPARC to data from the MSG SEVIRI sensor was described in<br />

this report. Three major modifications were presented:<br />

1. In addition to F, SPAR<strong>CM</strong>SG outputs a <strong>cloud</strong> mask with three classes (clear,<br />

thin <strong>cloud</strong>, thick <strong>cloud</strong>) based on thresholds applied to F. The application <strong>of</strong><br />

thresholds to F facilitates the trans<strong>for</strong>mation <strong>of</strong> the <strong>cloud</strong> mask from a clear-<br />

sky conservative to a <strong>cloud</strong> conservative state. In addition, the <strong>cloud</strong> mask<br />

error <strong>for</strong> each pixel is calculated based on F, which may be a valuable source<br />

<strong>of</strong> in<strong>for</strong>mation <strong>for</strong> users <strong>of</strong> this simplified <strong>cloud</strong> mask.<br />

2. SPAR<strong>CM</strong>SG is a stand-alone <strong>algorithm</strong> and does not rely on auxiliary data.<br />

In contrast to the original SPARC version, which uses an external surface<br />

skin temperature dataset as a reference in the calculation <strong>of</strong> the T-score,<br />

SPAR<strong>CM</strong>SG takes advantage <strong>of</strong> the high temporal resolution <strong>of</strong> MSG SEVIRI<br />

to derive a brightness temperature mosaic from SEVIRI clear-sky observations<br />

obtained during a preceeding time interval.<br />

3. A new snow detection module is implemented. Based on two newly defined<br />

sub-scores, it outputs both a daily snow mask as well as a snow score (S-score),<br />

which may be regarded as a measure <strong>of</strong> snow contamination within a pixel.<br />

For the period from July 2004 to June 2005, SPAR<strong>CM</strong>SG output was compared to<br />

<strong>cloud</strong> cover in<strong>for</strong>mation from the Alpine Surface Radiation Budget (ASRB) Net-<br />

work in the Swiss Alps: Cloud Free Index (CFI) and Partial Cloud Amount (PCA).<br />

Results show a good agreement between CFI and F mainly <strong>for</strong> daytime observa-<br />

tions. However, results also point to an underestimation <strong>of</strong> nighttime <strong>cloud</strong> cover by<br />

SPAR<strong>CM</strong>SG. Adjustment <strong>of</strong> the <strong>of</strong>fset and scale factors used to calculate the single<br />

scores may be required to account <strong>for</strong> this deficiency.<br />

If only two SEVIRI spectral channels are fed into SPARC to simulate data from<br />

the Meteosat First Generation (MFG), a good agreement between CFI and F is<br />

obtained during daytime. Finally, application <strong>of</strong> SPARC at full disk level revealed an<br />

underestimation <strong>of</strong> <strong>cloud</strong> cover compared to the Climate Monitoring (<strong>CM</strong>)-Satellite<br />

Application Facility (<strong>SAF</strong>) <strong>cloud</strong> mask. Both <strong>cloud</strong> masks should next be compared<br />

over longer time periods to synoptic <strong>cloud</strong> observations to evaluate possible biases<br />

and spatiotemporal inconsistencies.


SPARC <strong>algorithm</strong> 44<br />

Overall, SPAR<strong>CM</strong>SG has proven to provide interesting new opportunities <strong>for</strong> <strong>cloud</strong><br />

detection using the MSG SEVIRI sensor. Results are particularly promising with<br />

regard to the application <strong>of</strong> SPARC to data <strong>of</strong> the MFG satellite series, since the<br />

<strong>algorithm</strong> was found capable <strong>of</strong> generating <strong>cloud</strong> in<strong>for</strong>amtion with only a limited<br />

amount <strong>of</strong> spectral in<strong>for</strong>mation from two SEVIRI channels.


SPARC <strong>algorithm</strong> 45<br />

Bibliography<br />

Doms, G., and U. Schaettler (2002), The nonhydrostatic limited-area model lm<br />

part i: Dynamics and numerics. scientific documentation, deutscher wetterdienst,<br />

<strong>of</strong>fenbach, germany, available online: http://www.cosmo-model.org.<br />

Dürr, B. (2009), Technical report <strong>of</strong> msg-based ir <strong>cloud</strong>-index processing chain, Tech.<br />

rep., Sunergy GmbH.<br />

Dürr, B., and R. Philipona (2004), Automatic <strong>cloud</strong> amount detection by sur-<br />

face longwave downward radiation measurements, J. Geophys. Res., 109 (D5),<br />

D05,201–.<br />

EUMETSAT <strong>CM</strong> <strong>SAF</strong> (2009), Algorithm theoretical basis document.<br />

cm-saf product cm-02, cm-08 and cm-14. <strong>cloud</strong> fraction, <strong>cloud</strong> type<br />

and <strong>cloud</strong> top parameter retrieval from seviri, Available online:<br />

http://www.cmsaf.eu/bvbw/generator/<strong>CM</strong><strong>SAF</strong>/Content/Publication/atbd pdf/<br />

<strong>SAF</strong> <strong>CM</strong> DWD ATBD CFC CTH CTO SEVIRI 1.0,templateId=raw,<br />

property=publicationFile.pdf/<strong>SAF</strong> <strong>CM</strong> DWD ATBD CFC CTH CTO SEVIRI<br />

1.pdf.<br />

Khlopenkov, K. V., and A. P. Trishchenko (2007), SPARC: New <strong>cloud</strong>, snow, and<br />

<strong>cloud</strong> shadow detection scheme <strong>for</strong> historical 1-km AVHRR data over Canada,<br />

Journal Of Atmospheric And Oceanic Technology, 24 (3), 322–343.<br />

Lillesand, T., R. Kiefer, and J. Chipman (2004), Remote Sensing and Image Inter-<br />

pretation, fifth ed., John Wiley & Sons, New York.<br />

Mannstein, H., H. Broesamle, C. Schillings, and F. Trieb (1999), Using a meteosat<br />

<strong>cloud</strong> index to model the per<strong>for</strong>mance <strong>of</strong> solar thermal power stations, in Eumetsat<br />

Conference, p. 239246, Eumetsat, Copenhagen, Denmark.<br />

Marty, C., and R. Philipona (2000), The clear-sky index to separate clear-sky from<br />

<strong>cloud</strong>y-sky situations in climate research, Geophys. Res. Lett., 27 (17), 2649–2652.<br />

Mesinger, F., and Coauthors (2006), North american regional reanalysis, Bulletin <strong>of</strong><br />

the American Meteorological Society, 87, 343360.


SPARC <strong>algorithm</strong> 46<br />

Philipona, R., C. Marty, and C. Fróhlich (1996), Measurement <strong>of</strong> the Longwave<br />

Radiation Budget in the Alps, in IRS96, A. DEEPAK Publishing.<br />

Salomonson, V. V., and I. Appel (2004), Estimating fractional snow cover from<br />

MODIS using the normalized difference snow index, Remote Sensing <strong>of</strong> Environ-<br />

ment, 89 (3), 351–360.<br />

Schulz, J., P. Albert, H.-D. Behr, D. Caprion, H. Deneke, S. Dewitte, B. Dürr,<br />

P. Fuchs, A. Gratzki, P. Hechler, R. Hollmann, S. Johnston, K.-G. Karlsson,<br />

T. Manninen, R. Müller, M. Reuter, A. Riihelä, R. Roebeling, N. Selbach, A. Tet-<br />

zlaff, W. Thomas, M. Werscheck, E. Wolters, and A. Zelenka (2009), Operational<br />

climate monitoring from space: the eumetsat satellite application facility on cli-<br />

mate monitoring (<strong>CM</strong>-<strong>SAF</strong>), Atmospheric Chemistry and Physics, 9 (5), 1687–<br />

1709, doi:10.5194/acp-9-1687-2009.<br />

Walther, A., A. Heidinger, A. Thoss, and R. Roebeling (2009), Report on comparison<br />

study <strong>for</strong> the EUMETSAT workshop, Ascona, Switzerland.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!