34
This article was downloaded by: [Umeå University Library] On: 18 November 2014, At: 10:58 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK International Journal of Remote Sensing Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/tres20 Cloud mapping with groundbased photogrammetric cameras G. Seiz a b , J. Shields c , U. Feister d , E. P. Baltsavias a & A. Gruen a a Institute of Geodesy and Photogrammetry , Swiss Federal Institute of Technology ETH , ETHHoenggerberg , 8093 Zürich, Switzerland b Federal Ortice of Meteorology and Climatology MeteoSwiss , Kraehbuehlstrasse 58, 8044 Zürich, Switzerland c Scripps Institution of Oceanography (SIO) , University of California , San Diego (UCSD) , 9500 Gilman Dr., La Jolla, California 920930701, USA d German Weather Service (DWD) , Meteorological Observatory Lindenberg , Am Observatorium 12, 15848 Lindenberg, Germany Published online: 16 May 2007. To cite this article: G. Seiz , J. Shields , U. Feister , E. P. Baltsavias & A. Gruen (2007) Cloud mapping with groundbased photogrammetric cameras, International Journal of Remote Sensing, 28:9, 2001-2032, DOI: 10.1080/01431160600641822 To link to this article: http://dx.doi.org/10.1080/01431160600641822 PLEASE SCROLL DOWN FOR ARTICLE Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and should be independently verified with primary sources of information. Taylor and Francis shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of the Content.

Cloud mapping with ground‐based photogrammetric cameras

  • Upload
    a

  • View
    218

  • Download
    3

Embed Size (px)

Citation preview

Page 1: Cloud mapping with ground‐based photogrammetric cameras

This article was downloaded by: [Umeå University Library]On: 18 November 2014, At: 10:58Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954 Registeredoffice: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

International Journal of RemoteSensingPublication details, including instructions for authors andsubscription information:http://www.tandfonline.com/loi/tres20

Cloud mapping with ground‐basedphotogrammetric camerasG. Seiz a b , J. Shields c , U. Feister d , E. P. Baltsavias a & A.Gruen aa Institute of Geodesy and Photogrammetry , Swiss FederalInstitute of Technology ETH , ETH‐Hoenggerberg , 8093 Zürich,Switzerlandb Federal Ortice of Meteorology and Climatology MeteoSwiss ,Kraehbuehlstrasse 58, 8044 Zürich, Switzerlandc Scripps Institution of Oceanography (SIO) , University ofCalifornia , San Diego (UCSD) , 9500 Gilman Dr., La Jolla,California 92093‐0701, USAd German Weather Service (DWD) , Meteorological ObservatoryLindenberg , Am Observatorium 12, 15848 Lindenberg, GermanyPublished online: 16 May 2007.

To cite this article: G. Seiz , J. Shields , U. Feister , E. P. Baltsavias & A. Gruen (2007) Cloudmapping with ground‐based photogrammetric cameras, International Journal of Remote Sensing,28:9, 2001-2032, DOI: 10.1080/01431160600641822

To link to this article: http://dx.doi.org/10.1080/01431160600641822

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the“Content”) contained in the publications on our platform. However, Taylor & Francis,our agents, and our licensors make no representations or warranties whatsoever as tothe accuracy, completeness, or suitability for any purpose of the Content. Any opinionsand views expressed in this publication are the opinions and views of the authors,and are not the views of or endorsed by Taylor & Francis. The accuracy of the Contentshould not be relied upon and should be independently verified with primary sourcesof information. Taylor and Francis shall not be liable for any losses, actions, claims,proceedings, demands, costs, expenses, damages, and other liabilities whatsoever orhowsoever caused arising directly or indirectly in connection with, in relation to or arisingout of the use of the Content.

Page 2: Cloud mapping with ground‐based photogrammetric cameras

This article may be used for research, teaching, and private study purposes. Anysubstantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &Conditions of access and use can be found at http://www.tandfonline.com/page/terms-and-conditions

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 3: Cloud mapping with ground‐based photogrammetric cameras

Cloud mapping with ground-based photogrammetric cameras

G. SEIZ*{{, J. SHIELDS§, U. FEISTER", E. P. BALTSAVIAS{ and A. GRUEN{{Institute of Geodesy and Photogrammetry, Swiss Federal Institute of Technology ETH,

ETH-Hoenggerberg, 8093 Zurich, Switzerland

{Federal Ortice of Meteorology and Climatology MeteoSwiss, Kraehbuehlstrasse 58,

8044 Zurich, Switzerland

§Scripps Institution of Oceanography (SIO), University of California, San Diego

(UCSD), 9500 Gilman Dr., La Jolla, California 92093-0701, USA

"German Weather Service (DWD), Meteorological Observatory Lindenberg, Am

Observatorium 12, 15848 Lindenberg, Germany

Ground-based digital imager systems in the visible and near infrared region of

the solar spectrum have the potential to nicely complement existing instruments

and observation networks of National Weather Services with very accurate, high

spatial and temporal resolution, 2D and 3D macroscopic cloud data such as

cloud amount, cloud-base height and 3D cloud-base motion. This paper discusses

two current approaches to ground-based cloud sensing: the prototype instrument

used at ETH/MeteoSwiss within Cloudmap and Cloudmap2 for stereoscopy

tests, and a Daylight Visible/NIR Whole Sky Imager (WSI) system developed

and fielded by the Scripps Institution of Oceanography (SIO). The article

includes descriptions of the radiometric and geometric calibration methods.

Cloud amount, cloud-base height and cloud-base motion results from two ETH/

MeteoSwiss measurement campaigns and from the operational WSI use at the

German Weather Service (DWD) are shown. Finally, a case study with

coincident satellite and ground data illustrates that ground-based digital imager

systems are an interesting technique to validate satellite-based cloud-top heights

and cloud-top motion winds of vertically thin clouds.

1. Introduction

Clouds play a pivotal role in the interaction between the Earth’s climate andanthropogenic inputs. The Intergovernmental Panel on Climate Change (IPCC) has

highlighted the large scientific uncertainty associated with the role of clouds either to

accelerate or mitigate effects associated with increases in greenhouse gases. For a

detailed assessment, accurate global measurements of the relative location,

distribution and character of clouds (which have a strong impact on both the total

incoming radiation at the surface and the reflected radiation above the cloud field)

are necessary as described in the rationales of the EU projects Cloudmap and

Cloudmap2 (Cloudmap 2001, Cloudmap2 2001). For acquiring global coverage,satellite-based methods are used, but they benefit from calibration and validation

with ground-based measurements. In addition, ground-based sky imagery can

provide high-resolution images with a much higher sampling rate of less than

1 minute per image than present day satellite-based imagers and conventional cloud

*Corresponding author. Email: [email protected]

International Journal of Remote Sensing

Vol. 28, No. 9, 10 May 2007, 2001–2032

International Journal of Remote SensingISSN 0143-1161 print/ISSN 1366-5901 online # 2007 Taylor & Francis

http://www.tandf.co.uk/journalsDOI: 10.1080/01431160600641822

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 4: Cloud mapping with ground‐based photogrammetric cameras

observations. Due to the high variability of cloud cover and cloud optical depth in

time, both transmitted solar radiation and infrared atmospheric as well as terrestrial

radiation can vary in short time scales. Therefore, small uncertainties in time

integrals and averages of cloud parameters over time would also require small

sampling time steps in acquiring the individual data (e.g. Feister et al. 2003).

Ground-based cloud observations can also provide valuable information needed to

interpret and analyse solar irradiance measurements in different spectral regions

(e.g. Feister and Gericke 1998, Vasaras et al. 2001).

However, systematic and accurate ground-based measurements of the 3D cloud

field at high spatial and temporal resolution are not generally available from the

existing observation networks. At most climate stations of the national networks,

cloud macroscopic properties – mainly cloud cover, cloud type(s) and cloud-base

height – are still visually observed. Only at airports and some specific measurement

sites are automated weather sensors, ceilometers and more sophisticated instruments

(e.g. cloud radar, raman lidar) used to determine some cloud parameters

automatically and continuously.

To date, cameras and photogrammetric methods are rarely used within the

ground-based observational networks. Recent developments in the digital camera

market, including lower prices and larger image formats, are leading at the National

Weather Services (e.g. MeteoSwiss, German Weather Service DWD) to a revival of

the idea of photogrammetric cloud observation station networks, which was already

in discussion more than 100 years ago during the International Cloud Year 1896/

1897 (Koppe 1896). Operational use of analog images was not practical for most

applications due to the enormous amount of time necessary to analyse a single time

step by first scanning stereo image pairs with a photogrammetric scanner and then

finding corresponding points in the images by manual measurements. The new

digital systems have the major advantage of reducing significantly processing time

down to minutes, which gives them the potential of deriving cloud parameters in

near-realtime. Using fisheye lenses to image the full sky down to the horizon or

near-horizon, they have the potential to provide far more accurate measurements of

cloud fraction and distribution than e.g. ceilometers, because they acquire data over

the full sky simultaneously. For stereo camera systems, automatic, faster and more

reliable matching methods to solve the correspondence problem are a second

advantage. Finally, there is a qualitative advantage of data from a ground-based

photogrammetric system: these data are easier to interpret visually for a forecaster

and to link with the current synoptic situation than the point measurements from

active remote sensing instruments like ceilometers, lidars and radars.

In this paper, we demonstrate how the data from ground-based digital imager

systems in the visible and near infrared region of the solar spectrum system would

nicely complement existing observations of macroscopic cloud parameters, with

high spatial and temporal resolution data in 2D and 3D. Thereby, two research

efforts are discussed in detail: (1) the ETH prototype stereoscopic system for

determining the 3D cloud-base geometry and motion, and (2) the Whole Sky Imager

(WSI) systems developed at Scripps Institution of Oceanography (SIO), operation-

ally working in several design versions at several measurement sites world-wide since

the 1980s. These WSI systems are currently used for high spatial and temporal

resolution determination of cloud amount and type, and can potentially be used in

the future for determination of cloud optical thickness and, in stereo configuration,

3D cloud-base geometry and motion.

2002 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 5: Cloud mapping with ground‐based photogrammetric cameras

After an overview of the ETH prototype system and the WSI system in §2, the

radiometric and geometric calibrations of the two systems are described in §3. These

calibrations are an important prerequisite for obtaining accurate cloud parameters.

In §4, the retrieval methodologies for cloud amount, cloud type, cloud-base height,

cloud-base motion and cloud optical thickness are explained. In §5, the results of

these cloud parameters from two ETH/MeteoSwiss measurement campaigns and

from the operational WSI use at DWD are shown.

2. Camera and system requirements for cloud observation

The choice of the camera and other sensor hardware is crucial and depends in large

part on the desired features of the final system. This section will discuss two

photogrammetric imaging systems. The ETH prototype system was designed as apreliminary research system that would provide adequate image quality over a

moderate field of view, in order to evaluate stereoscopy algorithm approaches. The

WSI systems were designed to respond to a number of applications; for this article,

our primary interest is in its ability to operationally determine the 2D cloud

distribution (cloud amount and type) over the full sky.

2.1 ETH research system

The ETH camera system (figure 1) consists of a color digital CCD camera, currently

a Fujifilm S1 Pro, connected to a laptop computer with precise time information

from a GPS receiver or radio clock. For our first prototype, which was used during

the MAP campaign in October 1999, the KODAK DCS460 camera was used. A

main reason for having switched to the Fujifilm S1 Pro camera for the

Figure 1. ETH camera system.

CLOUDMAP: New satellite cloud products for cirrus and contrails 2003

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 6: Cloud mapping with ground‐based photogrammetric cameras

measurements at Zurich-Kloten airport in April 2002 was the fact that the KODAK

DCS460 camera unfortunately did not fulfill the stability requirements (a small

movement of the chip occurred during the MAP measurement campaign and was

possible because the chip is only attached along one chip border). The camera is

protected within a heated box (1–2uC above air temperature to avoid condensation

on image sensor) that is mounted on a tripod. Three leveling screws at the bottom of

the camera housing allow precise horizontal adjustment of the camera image plane.

Approximate adjustment of the camera azimuth parallel to the baseline was done

via a small telescope as the different cameras are always within sight of each other. A

more precise relative adjustment was then done by taking simultaneous sample

images with both cameras, overlaying the two images in an image processing

software and determining any residual rotation angle. Attached to the camera box is

a moving sun occultor device to prevent blooming effects caused by the sun. A

Nikon 18 mm wide-angle lens with a nominal viewing angle of 100u was used. As the

system is a prototype for research campaigns, the mechanical details of the system

can still be improved. Currently, an extension is in development to close a cap over

the camera lens automatically if it starts to rain/snow. Under these conditions, no

measurements are possible with this system.

The CCD array of the KODAK DCS460 is a KAF-6300 with 307262048 pixels,

each 969 mm2, with a Bayer colour filter (Bayer 1976). In the KODAK image

processing software the 6 rows and columns around the edge of the array are

discarded. The RGB values of the remaining 306062036 pixels are calculated with

the KODAK proprietary Active Interpolation algorithm from the red, green and

blue filter values (Adams et al. 1998). The dark current noise of this sensor is quite

substantial (10pA/cm2 at 25uC) and especially influences the long exposure night

images used for exterior orientation determination with stars (see §3). Therefore,

images were taken at various exposure times between 0.002 and 240 seconds with the

lens cap closed for each camera to analyse the dark current noise. It was shown that

the dark current noise is camera-dependant, spatially variable, temporally stable and

increases with longer exposure times.

The Fujifilm S1 Pro CCD array is a SuperCCD with 3.4 million octagonal pixels

and a special color filter as described in Tamayama et al. (2000). In the proprietary

Fujifilm processing software the octagonal color array counts are interpolated to

304062016 square pixels with a pixel size of 7.5 mm. The new CCD design attempts

to increase sensitivity, dynamic range, signal-to-noise ratio and image quality. The

30 s night images show much less noise than the Kodak DCS460 images; stars can be

extracted directly from the 30 s images without any previous flat field subtraction.

To obtain exposure times of up to 150 s, the camera offers a feature of taking a

sequence of up to five images without storage delay (temporal storage array) that

can then be summed up and clipped above the maximum grey value by image

processing software.

2.2 Daytime VIS/NIR WSI

Digital fully automated Whole Sky Imagers were developed by Scripps Institution of

Oceanography (SIO) at the University of California, San Diego beginning in the

early 1980s. Seven of these Day WSI instruments were fielded in the mid- to late

1980s, acquiring data once a minute simultaneously at several sites, in order to

determine cloud distribution (Shields et al. 1989, Johnson et al. 1989, Shields et al.

1993a).

2004 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 7: Cloud mapping with ground‐based photogrammetric cameras

Following several years of experience with the Day WSI, a Day/Night Whole Sky

Imager was developed at SIO in the early 1990’s (Shields et al. 1993b). This system is

designed to acquire cloud and sky images under all conditions over a 24-hour

period, and to meet a variety of research needs (Shields et al. 1998a). A sample night

image of the Day/Night WSI is shown in figure 2.

More recently, a new Daylight Visible/NIR WSI was funded by DWD. It was

designed to have many of the features of the Day/Night WSI, but only operate

during the day and under somewhat more benign conditions (Feister et al. 2001,

Shields et al. 2003) (figure 3). This instrument is designed to acquire cloud amount

and distribution, as well as sky radiance distributions in up to 7 wavebands in the

visible and NIR. Like its predecessors, it is fully automated and environmentally

hardened. A sample image of the Daylight Visible/NIR WSI acquired at SIO is

shown in figure 4. Most of this discussion concerns the Daylight Visible/NIR WSI,

which will be designated the Daylight WSI or the current Day WSI for the

remainder of this article.

The original Day WSI used an 8-bit CCD, however it was necessary to acquire

two sets of images, offset by 0.5 log using neutral density filters, to provide onscale

data over the full sky under many conditions. One of the design criteria for the

current Day WSI is that the darkest part of the sky have no more readout noise than

1% of the signal, and a resolution of at least 1% under most conditions. Using the

Day/Night WSI daytime data acquired with a 16-bit CCD, we evaluated the desired

range, and found that a 12-bit CCD would be most appropriate for the new

Daytime Visible/NIR WSI. Further evaluations of readout noise, dark noise,

exposure mechanisms, chip size, and system linearity led to the choice of

Photometric’s Sensys 1600 camera. This camera is cooled to a stable temperature

Figure 2. A sample image acquired with the Day/Night WSI at night with no moon (18December 1997, site: SGP, exposure: 60000 s). Note the presence of transparent clouds in theimage.

CLOUDMAP: New satellite cloud products for cirrus and contrails 2005

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 8: Cloud mapping with ground‐based photogrammetric cameras

of 10uC, yielding a true 12-bit dynamic range and the stability required for

radiometric sensing.

The remainder of the optical system includes two filter wheels, a neutral density

filter, and a 180u field of view fisheye lens. A custom optical lens relay system

integrates these components and provides the appropriate image size and placement

for this camera system. The current Day WSI, like its predecessors, includes

environmental hardening. The camera and optics are housed in a nitrogen-purged

camera housing, and the camera housing and electronics are housed in an

environmental housing with an air conditioner/heater to protect the components. In

addition, a fan blows heated air onto the dome to remove condensing water or rain

droplets from the dome surface. One of the important features of the WSI is the

solar occultor that shades the front optics of the system, thereby minimizing stray

light and providing optimal image quality. This occultor is designed to shade not

just the portion of the lens which images the direct sunlight, but the full front optics.

In order to provide automation and maximize the flexibility for general research

the system may be controlled either manually or under automated computer control

via in-house designed electronics. Under automated control, an internal flux control

algorithm selects the exposure times in order to optimize the data quality in all

filters. Although the instrument is still in the early stages of development in

comparison with the Day/Night WSI, quite a bit of development of processing

capabilities has been done. In post-processing, the instrument can provide calibrated

radiance measured simultaneously in each of approximately 740,000 directions, as

well as providing irradiance over the full sky and irradiance and average radiance in

selected regions of interest. This can be done in up to 10 selected regions of interest.

The resolution in the center of the image corresponds to less than 0.15u per pixel and

increases to about 0.25u per pixel close to the horizon (figure 5).

Figure 3. The Daylight Visible/NIR Whole Sky Imager fielded at DWD’s Potsdam Site.

2006 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 9: Cloud mapping with ground‐based photogrammetric cameras

The radiance distribution can be measured in up to 7 wavebands. Figure 6(a)

shows both the spectral filter transmissions measured in the DWD and MPL labs

and the CCD normalized spectral responsivity. In figure 6(b), the net normalized

system spectral responsivities of the different channels (based on the product of the

(a)

(b)

Figure 4. Sample images acquired at 650 nm from the Daylight Visible/NIR WSI. Left: 9July 2002, 10:50 UTC, contrails and 5 octa Ci partly developed from contrails; right: 13September 2002, 13:50 UTC, 5 octa Cu and 2 octa Ci fib; as a result of strong forwardscattering of solar radiation by cloud particles, the clouds at angular positions closer to thesun’s position appear to be brightened.

CLOUDMAP: New satellite cloud products for cirrus and contrails 2007

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 10: Cloud mapping with ground‐based photogrammetric cameras

sensor responsivity and the transmittances of the spectral filters and neutral density

filter) are illustrated.

3. Camera calibration

The radiometric and geometric calibrations are an important prerequisite for

obtaining accurate cloud parameters. Radiometric calibration is necessary for cloud

amount, cloud type and cloud optical thickness retrieval, while geometric

calibration is needed for stereo cloud-base height and motion determination. The

WSI instrument is radiometrically and geometrically calibrated, while the ETH

system is only geometrically calibrated.

3.1 Radiometric calibration

The determination of the 2D cloud fields with the WSI does not currently utilize the

calibrated radiance images, however portions of the radiometric calibration results

are used in the 2D cloud algorithm and the absolute results have potential in further

cloud algorithm development. Perhaps the most important aspect of calibration is

designing an instrument that acquires appropriate data for calibration, i.e. with

thermal stabilization, stray light control, the ability to acquire dark images, and so

on. The design considerations and the calibrations are discussed in detail in Shields

et al. (2003); in this section, a brief overview of the radiometric calibrations is given.

The radiometric calibrations used with the Daylight WSI include the following:

N Dark field correction, to correct for the thermally generated electrons and the

electronic bias. Although dark calibrations are evaluated in the laboratory,

the actual dark images are acquired in the field and applied immediately to the

field data.

N Flat field correction, to correct for spatial non-uniformities in the CCD. These

are acquired using a 1 m integrating sphere. For the Daylight WSI, the flat field

deviation was insignificant, and no correction was made.

Figure 5. Spatial resolution of the Daylight Visible/NIR WSI based on its geometricalcalibration.

2008 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 11: Cloud mapping with ground‐based photogrammetric cameras

N Exposure corrections, to correct for shutter- and camera-timing effects. These

are acquired on a 3 m bench using calibrated lamps traceable to NIST, and a

calibrated reflectance plaque. The exposure correction is about a 10%

correction at exposures of 100 ms and 1% at exposures of 1000 ms.

(a)

(b)

Figure 6. (a) Spectral transmissions of the filters as well as normalized spectral responsivityof the CCD of the Daylight Visible/NIR WSI. (b) Normalized spectral responsivities of the 7Day Vis/NIR channels derived as spectral combinations of transmissions of broad-bandfilters, neutral density filters and camera responsivity.

CLOUDMAP: New satellite cloud products for cirrus and contrails 2009

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 12: Cloud mapping with ground‐based photogrammetric cameras

N Linearity corrections, to correct for radiometric sensor non-linearity.

Linearities are measured on the 3 m bench. Corrections are about 2% or less

in mid-range and about 5% at the extremes of the range, except in high gain

where it is larger.

N Absolute radiance calibration, which converts the values to absolute spectral

radiance in SI units. The absolute calibration data were acquired on the 3 m

bench at SIO using lamps calibrated to absolute standards at the Physikalisch-

Technische Bundesanstalt (PTB) in Germany. In order to process the absolute

calibration, it is also necessary to characterize the spectral characteristics of the

filters, the CCD and any other spectrally-significant components in the system.

N Rolloff corrections, which correct for the change in system sensitivity as a

function of angle from the normal. These changes result primarily from the

changes in solid angle per pixel, but can also be affected by other losses such as

vignetting within the lens.

The calibration accuracy depends on the calibration lamp, which typically has an

uncertainty of 1–3% depending on the wavelength region and hours of use since

calibration. The overall calibration accuracy depends largely on how much

redundancy is used. Measuring the performance in multiple ways and evaluating

it with multiple means enables one to minimize stray light and other error sources.

The Daylight WSI calibration uncertainties were estimated at between 2 and 4%. A

field calibration device, for enabling more convenient recalibration of fielded

instruments, has been developed for the Day/Night WSI but not yet adapted for the

Daylight WSI.

3.2 Geometric calibration

3.2.1 ETH system. The calibration and orientation of the ETH camera system is

described in detail in Seiz (2003). It includes (1) the determination of the interior

orientation in the laboratory, and (2) an on-site measurement of the camera location

with GPS and orientation with so-called ‘sky control points’.

The interior orientation parameters are determined with a close-range photo-

grammetric reference field of 4.26261.2 m with 77 signalized and 20 coded points

at ETH, Institute of Geodesy and Photogrammetry (IGP). For each camera, 15

images were taken: from 5 camera stations (left high, left low, center, right high,

right low) at three different roll angles (290u, 0u, + 90u). The parameters were

calculated simultaneously with camera orientation data and 3-D object point

coordinates, employing a self-calibrating bundle adjustment. Ten additional

parameters were used to model systematic errors (Brown 1971): three parameters

of interior orientation (focal length offset dc, principal point coordinate offsets dxp

and dyp, five parameters modelling radial and decentering lens distortion (radial

coefficients k1, k2, k3; decentering coefficients p1, p2) and two parameters for a

differential scale factor and a correction of the non-orthogonality of the image

coordinate axes (Beyer 1992). The differential scale factor, the non-orthogonality

factor and some of the lens distortion coefficients proved to be insignificant.

The exterior orientation of both cameras had to be determined at the

measurement locations. As the cameras were horizontally mounted and the field

of view only includes clouds and sky, the traditional methods with ground control

points could not been used. Instead of ground control points, ‘sky’ control points

had to be measured. Two independent sets of sky control points were established:

2010 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 13: Cloud mapping with ground‐based photogrammetric cameras

during daylight, an airplane equipped with DGPS was flying along a specific flight

pattern; during clear nights, stars could be seen in images with long exposure times

and be taken as sky control points. The coordinates of each camera station were

additionally measured with GPS.

With the first approach, a previously calculated flight pattern was flown by the

KingAir of the Swiss Army, to get artificial sky control points. The flight lines were

parallel to the baseline of the two cameras. The highest line was 4000 m and the

lowest line at 1000 m above ground. The lines were along the left and right border of

the images and along the middle. The DGPS antenna of the airplane was manually

measured in every image. From the exact acquisition time of the image, the position

of the point could be determined from the GPS calculations and used as a control

point. With a bundle adjustment (with fixed interior orientation parameters from

the testfield calibration and station coordinates from static GPS), the orientation

angles were estimated together for both cameras. The image residuals showed an

accuracy of ,3 pixels across-track, consistent with the poor manual measurement

accuracy (due to the oblique viewing angle, the recognition of the plane and even

more of the position of the antenna was very difficult), but an accuracy of 10–

15 pixels along-track which was caused by the error in the precise acquisition time

(at a mean velocity of the KingAir of 100m/s, 10 pixels correspond to a time error of

about 100 ms at the mean flight height which is about the accuracy of the laptop

time). The stability of the angles in the bundle adjustment was improved with tie

points on clouds near the edges of the image.

The second method, with star images, was used to determine the exterior

orientation angles only. Additionally, the inner orientation parameters (same 10

parameter set as described above) of the camera were estimated, which is promising

for longer measurement periods where the inner orientation of a theoretically stable

camera could eventually change over time. This second method is also more realistic

for an operational sky imager network where recalibration of the orientation angles is

probably necessary at regular time intervals. The sky images with long exposure times

were taken during clear nights. When the exposure time was longer than about one

minute, the paths of the brightest stars could be seen between the noise (figure 7).

Although the noise represents a sum of dark current noise and sky background (i.e.

atmospheric scatter light, etc.), it could in this case, with very low sky background, be

modeled as dark current noise image alone, taken with a similar exposure time.

The enhanced star paths images were then further processed by specialized

software (Schildknecht 1994) to identify the stars corresponding to the Position and

Proper Motion (PPM) star catalog. From the star positions, the orientation angles

and the interior orientation parameters of each camera were calculated using the

same software package. The optimal exposure time is based on two factors: the

Figure 7. Left: star path and dark current noise, center: dark current noise, right: differenceimage.

CLOUDMAP: New satellite cloud products for cirrus and contrails 2011

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 14: Cloud mapping with ground‐based photogrammetric cameras

detectability of the linear feature (star track) within noise pattern, which increases

with increasing exposure time, and the linear form of the star path, which decreases

with increasing exposure time. The star path is assumed to be a straight line by the

processing software which finds the central point of each path by a centroid

operator. Only these central points are used in the calculations. The accuracy of the

calculated photogrammetric angles v, Q and k from the star images is ¡400.

3.2.2 WSI system. The WSI team has used a different approach to the geometric

calibration issue. A number of lens characteristics were tested, and it was determined

that a simple mapping from object space to image space could be determined by

measurements of objects at known angles with respect to the lens normal. When the

system is focused for infinite distance, objects in object space are in reasonable focus

from cloud distances up to a few inches from the lens. Studies of near-field and far-

field effects showed that a calibration using accurate markers a few meters from the

lens should yield valid results. Markers were placed on the walls of a room, with

approximately 0.1u accuracy with respect to a plumb bob location. Images are

acquired in this room with the full system facing the 0u marker. The system is

mounted on a rotary table, so that additional measurements can be taken at known

orientations to provide redundancy in the measurements. The resulting pixel-

position-to-angle calibration is accurate to approximately 2 pixels.

The largest source of uncertainty in the angular orientation is normally the

leveling and north alignment of the instrument, which are within approximately 1u if

installed by experienced teams. For this reason, the solar occultor was fabricated

with a 4-log neutral density that enabled detecting the solar disk in the image. In

certain applications, particularly with the original Day WSI system, the position of

the sun on the image was compared with the computed position of the sun.

Comparison of these points from images taken throughout the day enables one to

determine the level and orientation errors and make corrections accordingly. A

higher level of accuracy was desired for the Day/Night WSI, and is based on images

of the star field. This technique was developed by our colleagues in the ARM

Program (Tim Tooman, personal communication) and has been further modified

for other applications. It provides results typically accurate to approximately

0.25 pixels. A hybrid of these two techniques could also be applied to the current

Day WSI to yield more accurate geometric calibrations.

4. Retrieval of cloud properties

4.1 Cloud amount and distribution

This section provides an overview of the cloud algorithms that identify the presence

of clouds in the WSI imagery. The early work in daylight cloud algorithm

development using the WSIs in the early 1980s revealed that whereas cloud edges are

a very useful diagnostic, there are often significant edge gradients within clouds, and

weaker edge gradients at cloud edges. Similarly, signal level was useful but not

sufficient, because cloud signals could be significantly darker than sky signals under

certain conditions. An algorithm was developed based on the red/blue ratio. We

found that for opaque clouds, the red/blue ratio was significantly higher than for

clear sky backgrounds, and a fixed threshold in the red/blue ratio was quite effective

in identifying opaque clouds (Shields et al. 1989).

Thin clouds, being optically transparent, have a spatially varying red/blue ratio

which tends to be higher than the clear sky red/blue ratio. Like the clear sky ratios,

2012 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 15: Cloud mapping with ground‐based photogrammetric cameras

these ratios depend on the scattering angle with respect to the sun and the angle with

respect to the horizon. For thin clouds, a daytime algorithm was developed which

characterizes the thin clouds as a perturbation from the clear sky red/blue ratio. This

algorithm first determines the background clear sky red/blue ratio of the sky, as a

function of site, solar angle and viewing angle, from measured clear sky data, and

then evaluates the field images with respect to these background ratios. The

FASCAT radiative transfer model developed at SIO (Hering and Johnson 1984) was

used to evaluate anticipated background ratio performance, however we found that

the most accurate results were obtained using the background ratios extracted from

the site-specific clear sky measurements. Examples of a cloud algorithm image from

the current Day WSI are shown in §5.1.

Perhaps the most difficult challenge in the development of the WSI cloud

algorithm is distinguishing between aerosol and thin clouds. The shape of the clear

sky red/blue ratio distribution is relatively constant, and the aerosol load can be

treated by using a time- and day- varying normalization constant that modifies the

ratio distribution. This worked quite well with the original Day WSI, however it

required some human interaction in the archival processing. For potential real-time

processing, we installed NIR filters in the systems. These not only have a higher

contrast between thin cloud and clear sky, but they also should enable distinction

between the larger thin cloud droplets, as opposed to smaller aerosol droplets.

Although this NIR data has not yet been incorporated into the realtime day

algorithm for the D/N WSI or the new cloud algorithms for the current Day WSI, it

has been incorporated into an archival processing algorithm for the D/N WSI under

the auspices of the ARM program (Shields et al. 1998b). A night-time cloud

algorithm has also been developed for the D/N WSI, but is beyond the scope of this

article.

4.2 Cloud-base height

Cloud-base height and 3D cloud-base motion can be measured with ground-based

digital cameras in stereo configuration. Allmen and Kegelmeyer (1996) analysed

data acquired with a pair of the original Day WSI’s mentioned earlier, which had

been fielded in New Mexico by the SIO group. Due to the large base length (i.e.

instrument separation) chosen and due to inaccuracies of the exterior orientation,

the matching of the stereo pairs was very difficult. Furthermore, the matching

process was much slower than today due to slower computer processors. To

evaluate multi-view cloud-base height determination, a series of experiments were

run by ETH/MeteoSwiss within Cloudmap and Cloudmap2, using the ETH camera

system designed specifically for the stereoscopy experiments.

4.2.1 ETH system. The choice of an appropriate base length for cloud mapping is

difficult because of the wide height range of clouds (up to 15 km). There is a trade-

off between an as-large-as-possible overlapping area, optimal matching conditions

and an appropriate base-to-height ratio for the specific cloud situation, as will now

be explained in detail.

The overlapping area of a stereo pair is calculated as (figure 8):

overlap~100% 1{bc

xh

� �ð1Þ

CLOUDMAP: New satellite cloud products for cirrus and contrails 2013

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 16: Cloud mapping with ground‐based photogrammetric cameras

with b: base length [m], x: total width dimension of CCD sensor in baseline

direction, c: focal length, h: cloud height (m above ground).

For the MAP setup in October 1999, with b5850 m, c518 mm, x527.54 mm,

overlapping starts above 280 m cloud height; at 3.5 km, the overlap area is 92% and

at 10km 97%. In addition, shorter base lengths ((1 km) have the advantage that

matching is easier, faster and more reliable and many of the appearance difference

problems reported in Allmen and Kegelmeyer (1996) are avoided. But there is also

an important argument for larger base lengths. The higher a cloud is situated the

larger should be the base length due to the height accuracy sz:

sz~h2

bcspx ð2Þ

with spx: parallax measurement accuracy (m)

For our base length of 850 meters, an assumed parallax measurement accuracy of

1 pixel (59 mm) and a cloud height of 3.5 km and 10 km above ground, we get a

height accuracy of 7.2 m and 58 m respectively.

A solution to this trade-off is the use of more than two cameras, with different

base lengths. This of course also improves the reliability of the matching, and if high

performance is required, faster matching methods can be applied, due to the further

geometric constraints (Seiz 2003).

For the cloud-base height retrieval, different image processing algorithms,

including image matching, were applied. At IGP, many such algorithms have been

developed over the last 20 years and have been employed in matching of various

objects. Our approach within Cloudmap was to first use and test existing methods

and try to adapt them where necessary to model or reduce special problems

encountered with clouds. The use of existing matching algorithms was aiming at

achieving high success rate, accuracy and reliability. Aspects of processing speed

were not considered at this stage, although this is clearly of importance if a system

has to be used operationally and process large datasets. In the following, the

processing steps of the stereo pairs are shortly described; a detailed explanation of

the applied methods can be found in Seiz (2003).

For matching, the RGB images were first converted to 8-bit greyscale images. To

facilitate matching, a Wallis filter (Wallis 1976) was used as preprocessing. Wallis is

Figure 8. Illustration of overlapping area (shaded area) from two cameras.

2014 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 17: Cloud mapping with ground‐based photogrammetric cameras

an adaptive, local filter, which has been extensively used at IGP in image

preprocessing for matching (see e.g. Baltsavias 1991). The result is a radiometric

equalization of the stereo images and a contrast enhancement, i.e. more texture (see

figure 9).

For point extraction, two different operators were used, the Forstner operator

(Forstner and Gulch 1987) and the Harris operator (Harris and Stephens 1988). The

advantage of the Forstner operator compared to the Harris operator is the

(a)

(b)

Figure 9. Altocumulus cloud (a) in the original image and (b) after the Wallis filtering.

CLOUDMAP: New satellite cloud products for cirrus and contrails 2015

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 18: Cloud mapping with ground‐based photogrammetric cameras

possibility to extract points only along edges with a certain angle to the epipolar line

(i.e. search space of a point P of image A within image B, given accurate geometric

calibration parameters; otherwise the search space is extended to a so-called

‘epipolar band’, with a certain tolerance from the true epipolar line).

For matching, the Multi-Photo Geometrically Constrained Matching (MPGC)

package developed at IGP (Baltsavias 1991) was applied, which is based on Least-

Squares-Matching (Gruen 1985). With the geometric constraints, the search space is

restricted along the epipolar lines, increasing thus the success rate of the matching

and reducing matching problems (see figure 10). Depending on the cloud height

range, 3 to 5 image pyramid levels were used for the hierarchical matching

procedure. In the current tests, a fixed set of matching parameters was used in all

pyramid levels. Based on previous experience and some tests using supervised

matching of manually selected features, the following parameters were chosen as

optimal: patch size of 15615 pixels (pyramid level 0 and 1), 13613 pixels (level 2

and 3) and 11611 pixels (level 4 and 5), use of grey levels for matching, use of three

geometric parameters (two shifts and one scale) and decreasing weight for the

geometric constraints towards the higher-resolution pyramid levels (i.e. search space

is less restricted to epipolar line, but only to a so-called ‘epipolar band’).

For blunder detection, the MPGC software provides several statistical measures

that can be used to detect and exclude gross errors, e.g. cross-correlation coefficient,

a posteriori variance of unit weight from the least-squares adjustment, number of

iterations, etc. None of these measures can safely detect all blunders without

excluding good points. A combination of these quality measures can provide a

better diagnosis. More details about these measures and their use are given in

Baltsavias (1991) and Baltsavias and Stallmann (1993). The tests resulted in a

rejection of a certain percentage of the match points, usually about 5% to 25%,

depending on the matching problems and the selection of the thresholds. The quality

control tests were applied after matching in each pyramid level to avoid propagation

of wrong results to the next level.

4.2.2 WSI stereo feasibility. Although the WSI systems have not recently been

used for stereoscopic determination of cloud-base height, this section discusses

general considerations in converting them to yield this additional data product.

Figure 10. Illustration of geometrically constrained matching. Left: template (green); right:initial position of patch (green) and final solution of patch position (white). The search patchis moving along the epipolar line (blue line).

2016 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 19: Cloud mapping with ground‐based photogrammetric cameras

There are several important considerations in designing an imager specifically for

stereoscopy research and for later routine stereoscopy field measurements. As a

basic requirement, any set of two imagers should be synchronized. With the WSIs

discussed above, this is achieved via a GPS card on the system controller. The

question of angular resolution is important. A good angular resolution yields a more

precise height resolution in the stereoscopy. This can be achieved with higher

resolution cameras, but this results in more data volume, higher camera (and

potentially computer) costs, and larger processing time. If a full hemisphere field of

view is desired, this has implications both in the engineering of the sensor, and in the

reduced angular resolution. The final choice of angular resolution also directly

affects the flux levels on the CCD, and must be taken into account.

Although the current-generation WSIs have not been used for stereoscopy yet,

most if not all of the above preprocessing and matching techniques should be

applicable to the WSI data. For research into stereoscopy, the WSI has several

additional advantages. The images are acquired in several spectral filters. The NIR

data should be especially useful, because the contrast between sky and clouds is

larger, particularly under loaded aerosol conditions such as may often be found in

continental regimes. The instruments can be used to test the usefulness of calibrated

radiance in the stereoscopy algorithms. The D/N WSI can acquire images both day

and night, for monitoring of sky conditions under all conditions. The cloud

algorithms for cloud detection already exist for this instrument, and are continuing

to be improved. Data quality is very high, with high Signal/Noise ratios. The WSI is

especially suited for research applications requiring both the cloud distribution and

the absolute radiance distribution.

Whereas the large field of view is an advantage for cloud fraction determination,

it is a disadvantage for stereoscopy. Because the field of view and angular resolution

have been optimized for full hemisphere viewing, rather than stereoscopy, the height

resolution is (at least initially) poorer than indicated in §4.2.1. The viewing angle is

given by tan a5h/d, where h is cloud height and d is projected distance on the

ground between the cloud and the instrument. The computed height uncertainty for

the Daylight WSI, with its 180u field of view, is approximately 40 m near the zenith

at a height of 3.5 km, and 320 m at a height of 10 km, with a base length of 850 m

and roughly double these values with the current generation D/N WSI. However,

when the impact of uncertainty in the angular calibration is taken into

consideration, the very accurate angular calibration of the D/N WSI could more

than make up for this difference in the final analysis. In the final analysis, the most

accurate 3D assessment could probably be achieved by a full field of view WSI

which controls a smaller field of view peripheral instrument. The full field would

achieve the best 2D cloud distribution, while the smaller field of view sensor would

achieve the best height resolution. As a result of research using the WSIs as research

instruments, choices could be made regarding the best wavebands to use, and which

other features are most important, and a version of the WSI optimized for the

application could be designed.

4.3 Cloud-base motion (3D)

The same algorithms described in §4.2. can also be used for deriving 3D cloud

motion by matching corresponding features in the stereo image sequences. Various

time intervals between 30s and 120s were tested during MAP to evaluate the limits

for the matching/tracking algorithm. The tests showed that 120s can be too long in

CLOUDMAP: New satellite cloud products for cirrus and contrails 2017

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 20: Cloud mapping with ground‐based photogrammetric cameras

many cases, as there is not only a large displacement of the clouds given a high cloud

motion speed (which requires good approximations for the tracking), but more

problematic, significant shape changes of the cloud structures (see figure 11).

A further constraint is the amount of displacement, especially for lower clouds, so

that the overlapping area of the subsequent images gets too small. The amount of

displacement (in pixels) can be calculated as:

dx, dyð Þ~Dt u, vð Þ c

h

nr pixels x

xð3Þ

with (u, v): motion speed in x- and y-direction, x: dimension of sensor in baselinedirection, c: focal length, h: cloud height (m above ground).

In reality, cloud motion is usually larger at higher cloud heights so that for usual

motion amounts of 10–20 m/s, the displacement within 30s should be small enough

for still getting a large overlapping area for 3D cloud-base motion retrieval.

4.4 Cloud optical thickness

The WSIs also measure the sky radiance distribution (absolute values of radiantpower per unit area and per unit solid angle) with high spatial and temporal

resolution in several spectral bands. Sky radiance contains information on the

Figure 11. Problems that arise in image tracking if the time difference between twoconsecutive stereo pairs is too large. Top: stereo pair at time t0; bottom: stereo pair at time t1.There is a 2-minute time difference for these images obtained on 13 October 1999.

2018 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 21: Cloud mapping with ground‐based photogrammetric cameras

atmospheric extinction and may, in the future – in combination with the cloud

distribution – be used to derive estimates of optical features of the atmosphere such

as cloud optical thickness.

5. Results

In the following, results from the operational WSI measurements at DWD-Potsdam

and from two ETH research measurement campaigns, within the Mesoscale Alpine

Programme (MAP) in October 1999 and at Zurich-Kloten airport in April 2002, are

shown.

5.1 Whole sky imager (WSI) at the German weather service (DWD)

A Daylight Visible/NIR WSI with seven spectral channels was installed at the DWD

Meteorological Observatory Potsdam (52u229 E, 13u59 E, 107 m asl) in December

1999 and has been continuously working since that date (Feister et al. 2001). The

instrument worked reliably, with only two interruptions of operation in the summer

of 2000 due to an incorrect alignment of filters in the filter wheels and due to a

failure of the air conditioner. Dew and ice on the dome at winter time periods have

been removed by an additional fan blowing hot air from inside the environment

housing on the dome, though it has not been capable of removing ice that built at

times from rain droplets on the occultor arm with air temperatures below 0uC. The

daily schedule has been chosen as a compromise between sufficient sampling rate

and storage capacity with sequences of 7 images taken every 10 minutes between

sunrise and sunset. Shorter sampling rates of 1 sequence taken every two minutes

were chosen for selected time periods, when comparisons with other instruments

were performed. In an uncompressed mode, one image has a size of about 1.9

Mbytes. With an average amount of more than 1 GB of data per day, more than 1.5

Tbytes of dark corrected images have been collected over four years of operation.

While radiance files are even bigger with about 7.5 MB for each of them, the

processed files that contain the cloud cover for optically thick and optically thin

clouds are much smaller than the image files (a few kB per day only) and can be

processed quite quickly. An example of a processed cloud decision image for

optically thick clouds is shown in false colors in figure 12. The total cloud fraction

derived from this image is 0.36, with 0.21 of opaque and 0.15 of thin cloud fraction.

Conventional cloud observations provided a total cloud fraction of 0.25.

5.1.1 Conventional cloud observations. Ground-based cloud observations are

performed in meteorological networks according to guidelines of the World

Meteorological Organization (WMO 1975–1987). Clouds are classified in 4 different

categories as cloud families, i.e. low clouds (Stratocumulus and Stratus), medium–

level clouds (Altocumulus and Altostratus), high-level clouds (Cirrus, Cirrocumulus

and Cirrostratus) and clouds with vertical extension (Nimbostratus, Cumulus and

Cumulonimbus). The origin of that classification dates back to the English

manufacturing chemist and pharmacist Luke Howard (1772–1864). In the modern

synoptic meteorological code, only 3 categories are in use, i.e. Ns is assigned to

medium-level clouds, and Cu and Cb are considered part of low-level clouds. Each

of the three categories can have one out of nine different states and one out of nine

different cloud covers ranging from 0 octa (cloudless) to 8 octa (overcast). The digit

9 for cloud cover is used if clouds in the respective layer cannot be recognized due to

masking clouds below that layer, or in the case of fog at the ground (visibility less

CLOUDMAP: New satellite cloud products for cirrus and contrails 2019

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 22: Cloud mapping with ground‐based photogrammetric cameras

than 1 km). To estimate cloud cover, the observer divides the sky in 8 sub-areas of

equal size and counts the cloud-covered parts as ‘octa’ or ‘eighths’, or also called

‘octaves’. It is recommended in the manual of cloud observations that the total sky

should be observed ‘quasi-continuously to recognize the development and move-

ment of clouds and of the cloud cover at different levels over time’. Even if observers

are highly motivated and experienced, many other tasks to be performed at a

modern weather station will probably prevent an observer from completely meeting

that recommendation. Furthermore, the observer has to decide on one hourly value

of cloud type and cloud cover for each of the three categories, even though there can

be a high degree of variability of cloud cover within one hour of time. Therefore, it is

common to report those values that are observed about 10 minutes before the full

hour and assign those values to the full hour when the reports are to be sent to the

respective data center. A cloud cover and cloud type reported and sent from the

synoptic station to a weather center for the time 12 UTC may thus be more

representative for 11:50 UTC. In that respect, the cloud cover estimated by an

observer does not need to be a good estimate of the hourly average of cloud cover,

particularly, if atmospheric conditions are changing fast. In addition to the

uncertainty due to sampling frequency, the observer’s skill in estimating cloud cover

must be added to the total uncertainty of conventional cloud observations. The

Figure 12. WSI cloud decision image at Potsdam on 7 April 2002 at 11:30 UTC; total cloudfraction from the WSI cloud decision algorithm is 0.36, with 0.21 opaque (white and gray) and0.15 thin cloud fraction (yellow). Conventional cloud observations at the synoptic stationestimated a total cloud fraction of 0.25 Cu clouds.

2020 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 23: Cloud mapping with ground‐based photogrammetric cameras

uncertainty of total cloud cover estimated by a skilled observer at a weather station

is expected to not exceed ¡1 octa, but it may be slightly higher for clouds at higher

levels that are partly masked by cloud layers below that higher level. In addition, the

actual observation time may deviate by up to 5 minutes from the scheduledobservation time, i.e. in the example given above, it may vary between about 11:45

UTC and 11:50 UTC.

5.1.2 Preliminary comparison between WSI and cloud observations. An example of

cloud cover data derived from WSI red/blue image ratios is shown in figure 13. The

sampling rate of 2 minutes selected for that comparison shows much more detail in

short-time cloud cover variations than can be provided by conventional cloud

observations, which are collected every hour at the nearby weather station.

Although it can be expected from any visual observations that they are rather

subjective and dependent on the skill of the observer, our comparison showed a

good correspondence between the visual observations and the results from the skyimagers. Fast moving clouds would require an even higher sampling rate to reflect

the real cloud conditions at a time and not to erroneously bias hourly or daily

averages of cloud cover. Correspondingly, the sampling rate is also critical to other

parameters that can be derived from measured radiances and/or sky images. Shown

for comparison in figure 13 are values of cloud cover that were derived from images

acquired by another instrument type for ground-based daylight cloud imaging. That

instrument, a Total Sky Imager (TSI-440) manufactured by Yankee Environmental

Systems (YES 2000) was available at the DWD in Potsdam for a time period at theend of 2000 on loan from Aero Laser in Munich (Dieter Haaks, personal

communication). It uses a 368-bit color camera with a 3526288 pixel array. The

camera looks down at a slowly rotating spherical mirror that has a black strip to

occlude direct solar irradiance from the camera (but does not shade the mirror). The

Figure 13. Cloud cover (0 corresponds to cloudless and 1 corresponds to overcast sky) fromWSI images (red/blue ratios) for optically thick clouds and from TSI images (low clouds), andcloud cover observations of the Potsdam weather station (low clouds and clouds at medium-level heights), 27 October 2000.

CLOUDMAP: New satellite cloud products for cirrus and contrails 2021

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 24: Cloud mapping with ground‐based photogrammetric cameras

amount of degrees above the horizon for which the imager collects and processes

data can be selected by the user. For our comparison, it was set to 5u.Both WSI and TSI show similar short-time variations of cloudiness, with some

systematic lower cloud fractions observed by the TSI compared to the WSI, in

particular at low solar zenith angles (cloud cover from TSI is derived for solar

elevation angles larger than 3u only, while WSI data go down to solar elevation of

0u). Figure 13 is restricted to optically thick clouds, i.e. mainly those with low and

medium-height cloud-base height, because the WSI thin cloud algorithm for the Day

Imager was still under development at the SIO at the time of the comparison.

Some statistics of comparisons between total cloud cover derived from a set of

WSI images taken at Potsdam between May 2002 and April 2003, and from the

corresponding observations at the Potsdam weather station are shown in figure 14.

WSI total cloud cover values for the UTC minute 50 were rounded to octa and

compared with the cloud observations of the corresponding hourly cloud

observations. The spatial distance between the observer’s site and the imagers’ site

of 2 km will add to the observer’s uncertainty of ¡1 octa in estimating cloud cover.

It would translate into an angular distance of a cloud in the zenith between about

45u for a low cloud (CHB at 2 km) to an angular distance of 11u for a cloud at a

height of 10 km. In addition, the movement of clouds in connection with the time

uncertainty of the conventional cloud observation of about –5 … 0 minutes around

the observation time, which is ten minutes before the hour, would also add to the

uncertainty. If the clouds move at a speed of 10 km/h, this uncertainty in time

translates into a spatial difference of 0.8 km, which corresponds to an angular

distance of about 22u for low clouds and less than 5u for high clouds. A cloud

moving quite fast at a speed of 50 km/h would shift its location by about 4 km within

5 minutes or by an angular distance of about 64u for low clouds and 23u for high

clouds. A cloud movement of that size could significantly change the actual cloud

pattern. Depending on the direction of cloud movement in relation to the locations

Figure 14. Differences between total cloud cover in octas as derived from N53,817individual WSI measurements taken 10 minutes before the hour, and coincident hourly cloudobservations for the period May 2002 to April 2003 at Potsdam.

2022 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 25: Cloud mapping with ground‐based photogrammetric cameras

of both sites, this uncertainty due to horizontal movement could either enhance or

reduce the angular difference in the cloud position due to the spatial distance of

2 km between the observer and the imagers. However, as we have not compared

cloud patterns or cloud cover in different sky regions, but cloud cover referring to

the whole sky, both the imagers and the observer will generally see the same clouds

within their 180u fields of view, even though the observations are taken at slightly

different places and with small differences in time.

It can be seen in figure 14 that about 30% of the differences in cloud cover derived

from WSI images and observers are zero, 2/3 of the differences are within ¡1 octa,

and 4/5 of the differences are within ¡2 octa. Neither a systematic difference with

both completely independent types of observations nor a dependence on solar zenith

angle can be seen in the results of the comparison, except for very high solar zenith

angles (H.85u), when the cloud decision becomes more difficult due to scattering of

solar radiation from regions close to the sun position that are very bright and that

result in indeterminate red/blue WSI channel ratios in the cloud decision algorithm.

The comparison also shows that the typical occurrence of cloud cover at the

Potsdam site is quite well reflected by the present version of the WSI cloud decision

algorithm. Figure 15 shows the frequency distribution of cloud cover derived from

WSI data (first sample) and from cloud observations (second sample) for the same

dates and times. It can be seen in the figure that both sources of data show a similar

overall frequency pattern. Differences in the patterns are seen with cloudless

conditions, when the cloud observations show a much higher percentage of cases

than the WSI algorithm. The new WSI postprocessing algorithm can also separate

between optically thick and optically thin clouds. This feature will not be further

discussed here, but it allows us to study the causes of those differences in more

detail. With the thresholds used in the cloud decision algorithm in postprocessing

the WSI images, mostly very thin Cirrus clouds that show up in the images have not

been classified as thin clouds by the cloud decision algorithm. This misinterpretation

Figure 15. Occurrence of total cloud cover in octa for the period May 2002 to April 2003 atPotsdam. Cloud cover was derived from N53,817 WSI cloud decision images taken10 minutes before the hour, and coincident hourly cloud observations.

CLOUDMAP: New satellite cloud products for cirrus and contrails 2023

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 26: Cloud mapping with ground‐based photogrammetric cameras

could have been avoided by adjusting the thresholds of the clear sky reference values

before processing the images, but on the other hand, that approach could have

enhanced the number of misclassifications for haze conditions. The problem is

expected to be solved by the latest version of the cloud decision that is under

development at the SIO.

5.2 ETH research campaigns

The data acquisition for the ETH system data analysed in this section took place

during two ETH/MeteoSwiss measurement campaigns, (1) within the Special

Observation Period (SOP) of the Mesoscale Alpine Programme (MAP) at Mels,

Switzerland, in October 1999 (MAP Science Plan 1998), and (2) at Zurich-Kloten

airport, Switzerland, in April 2002.

During MAP, our two camera locations were separated by 850 m horizontally and

were visible from each other; the baseline direction was parallel to the valley, from

NW to SE. The choice of an appropriate base length for cloud mapping with only

two cameras is a difficult trade-off, as described in §4.2. For the second campaign,

the airport of Zurich-Kloten was chosen as the measurement site is equipped with

Vaisala 25K ceilometers which allowed an accurate comparison of cloud-base

heights with the results from our system. In §5.2.1 and §5.2.2, the use of radiosondes

and ceilometers as comparison instruments for cloud-base height is shortly

explained; a more detailed discussion can be found in Seiz (2003).

5.2.1 Radiosondes. During the MAP SOP in autumn 1999, several temporary

radiosonde stations were set up in the three target areas. They consisted of two types

of sondes: low-level sondes, measuring temperature and wind, and high-level sondes,

measuring pressure, temperature, humidity and wind. The five MAP low-level

sounding stations did not allow a reliable cloud boundary retrieval from the

temperature and wind profiles, so these soundings were not considered in our

comparison. The high-level sounding station of Diepoldsau (47.37 N, 9.66 E), which

was used for the comparison case of 08 October 1999, was equipped with ground

station P-760 receivers and used the same SRS-400 sonde as the operational

soundings from Payerne (Richner 1999). While the operational version of the SRS-

400 measures the dewpoint temperature with a dewpoint mirror for high precision

humidity values, the MAP SRS-400 sondes were equipped with a carbon element for

relative humidity measurement. Consequently, the accuracy of the SRS-400

sounding data was ¡2 hPa for pressure, ¡0.3 K for temperature and ¡2% for

relative humidity.

Radiosondes do not directly measure the vertical distribution of clouds, so that

cloud layers have to be inferred from their temperature and humidity profiles. A

common visual method to detect cloud layers is to plot the temperature and

dewpoint temperature profiles (see figure 16). At height levels where there is a small

difference between the two temperatures (visible in the plots at heights where the

two profiles approach each other) the atmosphere is saturated and cloud occurrence

is likely. However, it is important to remark that such saturated, or nearly saturated,

layers are no guarantee for the actual presence of clouds. Given the temperature and

humidity sensor accuracies from above, the accuracy of cloud-base heights derived

from radiosondes is about 200–300 m, until a height of about 8 km. Above 8 km, the

humidity measurements are too inaccurate due to the used SRS-400 sonde and due

to the low absolute humidity values at these heights.

2024 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 27: Cloud mapping with ground‐based photogrammetric cameras

5.2.2 Ceilometers. Today, operational ceilometers are installed at most airports

around the world. A ceilometer measures the height of cloud base above ground

level. It transmits laser pulses vertically into the atmosphere and measures the time

delay for the pulses to return to the ceilometer after reflection by small water

droplets in the clouds. The Vaisala 25K ceilometer (Vaisala 2003) installed at

Zuerich-Kloten airport uses a pulsed diode, InGaAs MOCVD laser at a wavelength

of 905 nm that transmits approximately 59000 light pulses vertically towards the

clouds each second. As the backscatter signal of a single pulse is weak, signals are

summed over intervals of 15 s to 30 s to improve signal-to-noise ratios. Instrument

software processes these signals into cloud-base height values by applying a

threshold on the return signal to distinguish between pure noise and noise summed

with a cloud signal. The resolution of the Vaisala 25K ceilometer is 50 ft (i.e. about

16 m) with an accuracy of ¡2% of the cloud height. The maximum detection height

of the Vaisala 25K ceilometer is around 259000 ft or 8 km.

5.2.3 Cloud-base height. Table 1 shows the result of two selected MAP stereo

pairs: an altocumulus and a cirrus situation. The cases were chosen with respect to

the MAP validation data available. For the 13 October 1999 case, the DSM points

were manually grouped into the two distinct cloud layers to enable comparison to

available validation data. The automation of cloud layer (or object) segmentation as

well as the 3D modeling and cloud visualization are important tasks which will be

further investigated within Cloudmap2.

To get a quantitative evaluation of the matching quality, the computed DSM has

to be compared with some reference data. As no high accuracy ceilometer was

available in the MAP composite observing network, the best validation of the results

Figure 16. Sounding from Diepoldsau, launched on 08 October 1999, at 11:00 (temperature;dewpoint). A strong inversion is visible around 4000m above sea level, indicating a cloud layerwith a base height of 3900–4000 m.

CLOUDMAP: New satellite cloud products for cirrus and contrails 2025

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 28: Cloud mapping with ground‐based photogrammetric cameras

is the comparison with semi-automatic matched points (i.e. manual location of

approximate value in patch image). Table 2 shows the results of the comparison with

about 60 semi-automatically measured points within the overlapping area of the

stereo pairs. Two versions were calculated for each case, one with the original

images and a second version with Wallis filtered images. As could be expected, the

preprocessed version performed better. For the 08 October 1999 images, with a

mean cloud height of 4000m, the expected z-accuracy according to Equation 2 is

about 8 m; for the 13 October 1999 case, with a mean cloud height of 10000 m, the

expected z-accuracy is about 58 m.

The comparison with other MAP data was quite difficult as most of the other

systems were also research instruments and not operating continuously. The best

comparison data for low and middle clouds (,8 km) were the frequent (every 3h)

Diepoldsau radiosonde data (§5.2.1.), launched only a few km from the camera

locations. Figure 16 shows the MAP sounding of 08 October 1999, launched at

11:00. The extracted cloud-base height and wind data from the sounding are

summarized in table 3.

Table 1. Cloud-base height results from the ground-based imager system for two selectedMAP cases (heights above sea-level).

Date, time Cloud typeMean cloudheight (km)

Cloud heightrange (km)

8 October 1999, 10:58 Altocumulus 4.0 3.8–4.213 October 1999, 10:16 Cirrus, two layers 8.0 7.8–8.1

10.9 10.7–11.1

Table 2. Statistical measures of the differences between the DSM results from the cloud-adapted MPGC and the manual measurements (positive: CBH from automatic

measurements.CBH from manual measurements).

ImagesNumber of

points Average (m) RMS (m) Max (m)

8 October 1999, original images 62 21.77 9.53 247.868 October 1999, preprocessed images 62 21.31 8.69 232.6413 October 1999, original images 47 17.0 98.9 563.813 October 1999, preprocessed images 47 15.6 60.1 413.6

Table 3. Comparison data for the two MAP cases: cloud-base height (CBH) and horizontalwind from radiosonde launched at Diepoldsau, about 30 min after camera measurements;satellite-based stereo cloud-top height (CTH) from ATSR-2 and cloud-top motion wind from

Meteosat-6 5min Rapid Scans (only available during second case).

DateRadiosondeCBH (km)

Radiosondehoriz. winddirection (u)

Radiosondehoriz. windspeed (m/s)

ATSR-2CTH(km)

M6 CMWdirection

(u)M6 CMWspeed (m/s)

08 October1999

3.9 296 8.5 – – –

13 October1999

8.5 280 18.8 11.0 275 19.3

–* 275 27.4

*No higher dewpoint measurements (only up to about 9–10 km)

2026 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 29: Cloud mapping with ground‐based photogrammetric cameras

In figure 17, the 3D cloud field results of 08/10/1999 is presented, to show the

height distribution over the whole overlapping area of the two cameras.

For vertically thin clouds (13 October 1999), the additional use of satellite data

was helpful. The cloud-top height and motion was extracted from the satellite data

of ERS2-ATSR-2 and Meteosat-6 according to the methods described in Seiz(2003).

In the following paragraph, one case study from the Zurich-Kloten airport

measurements in April 2002 is presented, to show a comparison of our stereo cloud-

base height retrievals with coincident ceilometer data. For the quantitative

(a)

(b)

Figure 17. Cloud-base height result from stereo images of 8 October 1999. Left: orthophoto;right: cloud-base height (CBH) field.

CLOUDMAP: New satellite cloud products for cirrus and contrails 2027

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 30: Cloud mapping with ground‐based photogrammetric cameras

comparison, seven time intervals with a spacing of 5 min were analysed between

10:30 UTC and 11:00 UTC on 12 April 2002 (table 4). The nearest 10 cloud points to

the location of the runway (RW) ceilometer were extracted from the stereo cloud-

base height results and averaged. Except for the 10:55 UTC time interval, all points

were from the same cloud layer, while at 10:55 UTC the points were grouped into

two distinct layers. For the 15 s ceilometer data the values within a small window of

¡60 s (i.e. 9 values) were averaged. As the results in table 4 illustrate, the stereo-

derived CBHs are quite consistent with the ceilometer values. However, we can see

that the ceilometer had problems with broken clouds when clear sky was reported.

The cameras with their larger FOV have a clear advantage for such broken cloud

cases.

5.2.4 Cloud-base motion. Table 5 shows the extracted motion parameters direction

and speed for (a) the 08 October 1999 dataset, with 30s interval and (b) the 13

October 1999 data set, with 120s interval. Although there are many blunders in the

13 October 1999 results due to the reasons mentioned above, the two distinct cloud

layers already identified in §5.2.1 with the cloud-base heights, are also clearly visible

here, with significant motion speed differences between the two layers, but about the

same motion direction.

The validation of cloud-base motion is even more difficult than the validation of

cloud-base height, as most of the conventional instruments measure the wind speed/

direction of the air and not of the clouds. In some cases, the cloud motion can

correspond to the wind at the cloud base, but in most cases, the comparison can be

misleading (e.g. orographic foehn clouds). Therefore, for the vertically thin cirrus

case (13 October 1999), the use of satellite data was again helpful. The horizontal

Table 4. Comparison of stereo camera results with ceilometer time series data, 19 April 2002,10:30–11:00 UTC.

Time (UTC)Camera system (nearest

10 points) (m ASL)Ceilometer (2 min average)

(m ASL)

10:30:00 2519¡9 2375¡4210:35:00 2123¡114 2198¡20510:40:00 2001¡3 1957¡1610:45:00 2168¡88 2080¡5610:50:00 2167¡114 2181¡18110:55:00 2211¡63 2106¡33

3877¡5011:00:00 5698¡114 – (clear sky)

Table 5. Cloud-base motion wind results from the ETH ground-based imager system for twoselected MAP cases.

Date, timeFrequency

(s)Horizontal motionwind direction (u)

Horizontalmotion windspeed (ms21)

Vertical motionwind speed

(ms21)

8 October 1999,10:58

30 310 10.2 0.5

13 October 1999,10:16

120 274 17.8 –1.1

276 25.8 1.3

2028 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 31: Cloud mapping with ground‐based photogrammetric cameras

cloud-base motion is compared to the cloud-top motion extracted from the MAP

Meteosat-6 5min Rapid Scans (see table 3). For the vertical cloud-base motion, no

comparison data was available. However, the calculated values seem to be

reasonable.

6. Conclusions

The results from the three ground-based digital photogrammetric camera systems

show the potential of such systems to complement existing measurement networks

with highly valuable cloud data.

In the retrieval of cloud amount, the WSI is shown to be capable of providing

cloud cover data for the whole sky as well as up to 10 readily selectable sky portions

for both optically thick and thin clouds with a sampling rate of down to about 1–

2 minutes. The version of Daylight Visible/NIR WSI applied in the German

Weather Service has been continuously used to collect sky images and radiances

over a period of four years. Post-processing of the large amount of raw data will

provide a data base of cloud cover and sky radiances that can be used to validate

satellite-based cloud retrieval techniques as well as to develop methods for

extracting other atmospheric parameters from the measured radiances. Further

operational use of this instrument at the DWD Radiation Center and studies using

its data can also be helpful in the design of simpler instruments that are suitable to

provide synoptic and climatological cloud data at networks of automatic weather

stations.

The stereoscopy tests with the ETH systems proved the high accuracy potential of

such a stereo camera system in the cloud-base height and motion measurement.

With the current system based on simple off-the-shelf components, the achieved

height accuracy is already about 0.5–1% of the cloud-base height (depending on the

matching accuracy for different cloud types). The system proved furthermore its

usefulness for comparison with satellite-based cloud products for vertically thin

cloud situations.

Finally, it has to be noted that ground-based sky imaging in the visible and near

infrared region bears the potential to derive other atmospheric parameters in

addition to the ones discussed in this paper. Among them are macroscopic and

optical cloud features such as cloud types and cloud optical depths, and even

microphysical cloud features such as droplet distributions as well as atmospheric

aerosol information, e.g. its thickness and size distribution.

7. Outlook

Improvements in the systems may concentrate on different aspects that depend on

the planned application of the specific instrument type. While a radiation and

climate monitoring station would probably need as many atmospheric parameters as

possible to be derived from the sky images and radiances (with low uncertainties,

high spatial resolution and high temporal resolution), a subset of parameters such as

total cloud cover, cloud-base height and cloud structures may be sufficient for

unattended field weather stations, which could be derived from much simpler and

therefore more cost-effective instrument versions. In any case, all the stations

between those two extremes mentioned that perform regular cloud observations

should be equipped by fully automatic and reliable instruments that keep instrument

system failures and the need of manual data processing at a minimum. Therefore, a

CLOUDMAP: New satellite cloud products for cirrus and contrails 2029

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 32: Cloud mapping with ground‐based photogrammetric cameras

combination of the WSI operability with the ETH system’s processing software

would be a good collaboration for operational stereo-photogrammetric retrieval of

cloud-base heights. Furthermore, in addition to improvements in the imaging

systems themselves, there is a need to develop software tools that make processed

data easily available to the potential user without time delay.

Considering these operational aspects, the WSI instruments would then be useful

for upgrading the existing prototype camera networks of the National Weather

Services to additionally measure quantitative cloud parameters operationally, such

as cloud amount, cloud height and cloud motion.

Acknowledgements

The MAP data was provided by the various principal investigators of the MAP

Rhine Valley target area and the Swiss Meteorological Institute. We thank Kodak

and FHBB Muttenz for providing the DCS460 cameras, the Swiss Army for the

calibration flights and the Astronomical Institute Bern (Martin Ploner, Thomas

Schildknecht) for the star processing. The work of ETH described in this paper is

funded by the Bundesamt fur Bildung und Wissenschaft (BBW) within the EU-

projects CLOUDMAP (BBW Nr. 97.0370) and CLOUDMAP2 (BBW Nr. 00.0355-

1).

We additionally thank Robert Fugate of Starfire Optical Range, Ann Slavin of

Boeing and Tim Tooman of Sandia National Labs for their technical advice and

feedback during the development of the Day/Night WSI. Other members of DWD

who were very helpful in the development of the Day Vis/NIR WSI were Klaus

Dehne and Rolf-Dieter Grewe. Andrew Beauto at Roper Scientific has been very

supportive in these efforts. We also thank the current members of the Atmospheric

Optics Group at SIO’s Marine Physical Lab who do development work on the

WSI’s and their algorithms: Richard Johnson, Monette Karr, Art Burden, Justin

Baker and Jerry Crum.

ReferencesADAMS JR., J.E., PARULSKI, K. and SPAULDING, K., 1998, Color Processing in Digital

Cameras. IEEE Micro, 18, pp. 20–30.

ALLMEN, M.C. and KEGELMEYER JR., W.Ph., 1996, The computation of cloud-base height

from paired whole-sky imaging cameras. Journal of Atmospheric and Oceanic

Technology, 13, pp. 97–113.

BALTSAVIAS, E.P., 1991, Multiphoto Geometrically Constrained Matching, Ph. D. disserta-

tion, Institute of Geodesy and Photogrammetry, ETH Zurich, Mitteilungen No. 49,

221 p.

BALTSAVIAS, E. and STALLMAN, D., 1993, SPOT stereo matching for DTM generation.

Proceedings of the Conference on Integrating Photogrammetric Techniques with

Scene Analysis and Machine Vision, 12–16 April, Orlando, USA. In Proceedings of

the International Society for Optical Engineering, 1944, pp. 152–163.

BAYER, B.E., 1976, Color imaging array, United States Patent 3,971,065, US Patent and

Trademark Office, Washington DC, USA.

BEYER, H.A., 1992, Geometric and radiometric analysis of a CCD-camera based

photogrammetric close-range system, PhD dissertation, Institute of Geodesy and

Photogrammetry, ETH Zurich, Mitteilungen No. 51, 186 p.

BROWN, D.C., 1971, Close-range camera calibration. Photogrammetric Engineering, 37, pp.

855–866.

CLOUDMAP, 2001, Project information available online at: http://www.ge.ucl.ac.uk/

research/cloudmap.

2030 G. Seiz et al.

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 33: Cloud mapping with ground‐based photogrammetric cameras

CLOUDMAP2, 2001, Project information available online at: http://www.ge.ucl.ac.uk/

research/cloudmap2 (accessed 15 January 2007).

FEISTER, U. and GERICKE, K., 1998, Cloud flagging of UV spectral irradiance measurements.

Atmospheric Research, 49, pp. 115–138.

FEISTER, U., SHIELDS, J., KARR, M., JOHNSON, R., DEHNE, K. and WOLDT, M., 2001,

Ground-based cloud images and sky radiances in the visible and near infrared region

from Whole Sky Imager measurements. CM-SAF Workshop, Dresden, 20–22

November 2000. EUMETSAT Proceedings, EUM P 31, pp. 79–88 (Darmstadt,

Germany: EUMETSAT).

FEISTER, U., KAIFEL, A., GREWE, R.-D., KAPTUR, J., REUTTER, O., WOHLFART, M. and

GERICKE, K., 2003, First performance results of two novel spectroradiometers

developed for fast scanning of solar spectral UV irradiance. SPIE International

Symposium on Optical Science and Technology, San Diego, CA, 3–8 August.

FORSTNER, W. and GULCH, E., 1987, A fast operator for detection and precise location of

distinct points, corners, and centers of circular features, Proceedings of the ISPRS

Intercommission Conference on Fast Processing of Photogrammetric Data,

Interlaken, Switzerland, 2–4 June, pp. 281–305.

GRUEN, A., 1985, Adaptive least squares correlation: a powerful image matching technique.

South African Journal of Photogrammetry, Remote Sensing and Cartography, 14, pp.

175–187.

HARRIS, C. and STEPHENS, M., 1988, A combined corner and edge detector. Proceedings of

4th Alvey Vision Conference, Manchester, UK, 31 August–2 September, pp. 147–151.

HERING, W.S. and JOHNSON, R.W., 1984, The FASCAT Model Performance Under

Fractional Cloud Conditions and Related Studies, University of California, San

Diego, Scripps Institution of Oceanography, Marine Physical Laboratory, SIO 85–7,

AFGL-TR-84-0168.

JOHNSON, R.W., HERING, W.S. and SHIELDS, J.E., 1989, Automated Visibility and Cloud

Cover Measurements with a Solid-State Imaging System, University of California,

San Diego, Scripps Institution of Oceanography, Marine Physical Laboratory, SIO

89-7, GL-TR-89-0061.

KOPPE, C., 1896, Photogrammetrie und Internationale Wolkenmessung. Braunschweig

Verlag, 108 p.

MAP IMPLEMENTATION PLAN, 1999, Available online at: http://www.map.ethz.ch/

SOPprep.htm.

RICHNER, H., 1999, Grundlagen aerologischer Messungen speziell mittels der Schweizer

Sonde SRS-400, Technical Report Nr. 61, Veroeffentlichungen der SMA-

MeteoSchweiz.

SCHILDKNECHT, T.H., 1994, Optical astrometry of fast moving objects using CCD detectors,

PhD Dissertation, Astronomical Institute, University of Berne, Geodatisch-geophy-

sikalische Arbeiten in der Schweiz, Vol. 49.

SEIZ, G., 2003, Ground- and satellite-based multi-view photogrammetric determination of 3D

cloud geometry, PhD thesis, Institute of Geodesy and Photogrammetry, ETH Zurich,

Switzerland, IGP Mitteilungen Nr. 80.

SHIELDS, J.E., KOEHLER, T.L. and JOHNSON, R.W., 1989, Whole Sky Imager, Proceedings of

the cloud Impacts on DOD Operations and Systems Workshop.

SHIELDS, J.E., JOHNSON, R.W. and KOEHLER, T.L., 1993a, Automated Whole Sky Imaging

Systems for Cloud Field Assessment, Proceedings of the Fourth Symposium on

Global Change Studies, 17–22 January 1993, published by the American

Meteorological Society, Boston MA.

SHIELDS, J.E., JOHNSON, R.W. and KARR, M.E., 1993b, Automated Whole Sky Imagers for

Continuous Day and Night Cloud Field Assessment, Proceedings of the Cloud

Impacts on DOD Operations and Systems Conference.

SHIELDS, J.E., JOHNSON, R.W., KARR, M.E. and WERTZ, J.L., 1998a, Automated Day/Night

Whole Sky Imagers for Field Assessment of Cloud Cover Distributions an Radiance

CLOUDMAP: New satellite cloud products for cirrus and contrails 2031

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014

Page 34: Cloud mapping with ground‐based photogrammetric cameras

Distributions, Proceedings of the 10th Symposium on Meteorological Observations

and Instrumentation, 11–16 January, published by the American Meteorological

Society, Boston MA.

SHIELDS, J.E., KARR, M.E., TOOMAN, T.P., SOWLE, D.H. and MOORE, S.T., 1998b, The

Whole Sky Imager – A Year of Progress, Proceedings of Eighth Atmospheric

Radiation Measurement (ARM) Science Team Meeting.

SHIELDS, J.E., JOHNSON, R.W., KARR, M.E., BURDEN, A.R. and BAKER, J.G., 2003, Daylight

Visible/NIR Whole Sky Imagers for Cloud and Radiance Monitoring in Support of

UV Research Programs, SPIE International Symposium on Optical Science and

Technology.

TAMAYAMA, H., SAITO, O. and INUIYA, M., 2000, High-definition still image processing

system using a new structure CCD sensor. SPIE Proceedings, 3965, pp. 431–436.

VAISALA, 2003, Available online at: http://www.vaisala.com.

VASARAS, A., BAIS, A.F., FEISTER, U. and ZEREFOS, C.S., 2001, Comparison of two methods

for cloud flagging of spectral UV measurements. Atmospheric Research, 57, pp. 31–42.

WALLIS, R., 1976, An approach to the space variant restoration and enhancement of images,

Proceedings of Symposium on Current Mathematical Problems in Image Science,

Naval Postgraduate School, Monterey CA, USA, November.

WMO, 1975–1987, International Cloud Atlas, Vol. I – Manual on the observations of clouds and

other meteors, World Meteorological Organization, revised edition, reprinted by

Selbstverlag des Deutschen Wetterdienstes, Offenbach, 1995.

YES, 2000, Total Sky Imager, Installation and User Guide, Yankee Environmental Systems,

Inc. Available online at: http://www.yesinc.com.

2032 CLOUDMAP: New satellite cloud products for cirrus and contrails

Dow

nloa

ded

by [

Um

eå U

nive

rsity

Lib

rary

] at

10:

58 1

8 N

ovem

ber

2014