scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Depolarization ratio profiling at several wavelengths in pure Saharan dust during SAMUM 2006.

TL;DR: In this paper, the vertical profiles of the linear particle depolarization ratio of pure dust clouds were measured during the Saharan Mineral Dust Experiment (SAMUM) at Ouarzazate, Morocco, close to source regions in May-June 2006, with four lidar systems at four wavelengths (355, 532, 710 and 1064 nm).
Abstract: Vertical profiles of the linear particle depolarization ratio of pure dust clouds were measured during the Saharan Mineral Dust Experiment (SAMUM) at Ouarzazate, Morocco (30.9 ◦ N, –6.9 ◦ E), close to source regions in May–June 2006, with four lidar systems at four wavelengths (355, 532, 710 and 1064 nm). The intercomparison of the lidar systems is accompanied by a discussion of the different calibration methods, including a new, advanced method, and a detailed error analysis. Over the whole SAMUM periode pure dust layers show a mean linear particle depolarization ratio at 532 nm of 0.31, in the range between 0.27 and 0.35, with a mean Angstr¨ om exponent (AE, 440–870 nm) of 0.18 (range 0.04–0.34) and still high mean linear particle depolarization ratio between 0.21 and 0.25 during periods with aerosol optical thickness less than 0.1, with a mean AE of 0.76 (range 0.65–1.00), which represents a negative correlation of the linear particle depolarization ratio with the AE. A slight decrease of the linear particle depolarization ratio with wavelength was found between 532 and 1064 nm from 0.31 ± 0.03 to 0.27 ± 0.04.

Summary (4 min read)

1. Introduction

  • Shape, size distribution and composition of aerosol particles influence their scattering characteristics and thus the radiative impact.
  • Thus, observations of the linear depolarization ratio at several wavelengths may be used in retrieval schemes (Dubovik et al., 2006) to improve the estimation of the microphysical properties of dust from optical measurements (D. Müller, personal communication, 2008; Wiegner et al., 2008).
  • A further technique is discussed, which improves the calibration of the depolarization channels.
  • The authors begin with the presentation of the polarization lidar technique (methodology, lidar setup) in Sections 2 and 3.

2. Polarization lidar method

  • Measurements of the linear depolarization ratio δ with lidars are often performed with the aim to just discern between the dry, the liquid and the ice phase of aerosols and clouds in the profiles of one lidar system, which requires only a relative measure of δ with a low accuracy of the absolute values.
  • Thus, the uncertainty of the absolute values must be known and should be small compared with the expected natural variance.
  • The total backscattered power P(r) with their dependence on the distance r from the lidar is described by the lidar equation P = ηβ(r)τ 2(r) r2 , (1) where η is the system constant, β the backscatter coefficient, and the factor τ 2 accounts for the atmospheric transmittance on the way from the lidar to the scattering volume, and back.
  • Furthermore the polarizing beamsplitter might be misaligned with respect to the plane of polarization of the emitted laser beam, and additionally, a rotation of the polarization plane is used for the relative calibration of the two receiver channels.
  • The total backscatter power P and the total backscatter coefficient β are the sum of both polarized components:.

2.1. Calibration

  • (11) With known δv in some range of the lidar signal, the authors can determine V∗ already from a regular measurement with eq. (11), which they call the ‘0◦-calibration’.
  • The linear depolarization ratio δm of air molecules is well known (Behrendt and Nakamura, 2002), and thus, the authors could use an aerosol-free lidar range in the free troposphere were δv = δm.
  • But already a very low amount of strong depolarizing aerosols, like dust or ice crystals in the assumed clean range, causes large errors of δv and thus of V∗. Furthermore, several instrumental uncertainties can add large errors.
  • The authors call this method the ‘45◦- calibration’.
  • If the polarizing filter has a negligible small transmission in the cross-polarized plane (k2 < 10−4), eqs. (12) and (13) are still valid, and the calibration errors are similar to those in Fig.

2.3. Retrieval of the linear particle depolarization ratio δp

  • The backscatter ratio R can be retrieved from the total signal P using, for example, the Fernald/Klett inversion with a reference value βp(r0) at a reference range r0 and known range-dependent lidar ratios S S = β P αp , (22) where αp is the particle extinction coefficient.
  • The values of δv and R are subject to systematic and statistical errors.

3. Polarization lidar in SAMUM

  • The linear depolarization ratio of the dust particles was measured during SAMUM with four lidar systems at four wavelengths:MULIS at 532 nm;POLIS at 355 nm (both from Munich);BERTHA at 710 nm and the airborne DLR HSRL at 532 and 1064 nm.
  • All three ground based systems can change the zenith angle of the sounding (scanning capability), which possibility was used frequently.
  • The authors assume that the profiles of the different lidar systems are well comparable as orographical effects of the surrounding are negligible and the dust layer was mostly sufficiently homogeneous over the considered periods.
  • For the DLR-HSRL and BERTHA, the details regarding the depolarization measurements complement the descriptions in the references mentioned below.

3.1. MULIS

  • Such a high signal exceeds the linear range of the data acquisition, especially that of the preamplifier, which would introduce signal distortions and errors in V∗, which are very difficult to assess.
  • To minimize diattenuation and polarization effects of the dichroic beamsplitters, all optical coatings are custom-designed (Laseroptik, Germany).
  • To reduce cross-talk, the custom-made PBC (Optarius, UK) has high transmittance Tp and high reflection Rs, and P‖ is detected in the reflected branch PR of the PBC (see eq. 15).
  • The accuracy of the ϕ = 0◦ position is better than ±0.2◦, and a ±45◦-calibration was performed every time the lidar setup had been changed.

3.2. POLIS

  • The portable lidar system POLIS is a small, rugged, two-channel lidar system with several modular detection options.
  • During SAMUM, it measured the parallel- and cross-polarized signals at 355 nm or the backscatter at 355 nm and the N2-Raman shifted wavelength at 387 nm.
  • The two different detector modules are rigidly mounted to the telescope by means of a superfinished circular flange with an angular scale, which allows the manual ±45◦-rotation of the whole depolarization detector module with respect to the laser polarization with an estimated accuracy of ±1◦ (Fig. 5).
  • Absorbing neutral density filters are used to adjust the sensitivity ranges of the individual detection channels.
  • Other technical details of this system can be found in Heese et al. (2002).

3.3. DLR-HSRL

  • A detailed description of the DLR-HSRL system and an assessment of its measurement accuracy can be found in Esselborn et al. (2008) and Esselborn et al. (2008).
  • Besides the direct extinction measurement at 532 nm by means of the HSRL technique, the system is designed to measure the linear depolarization ratio at 1064 and 532 nm.
  • A dichroic beamsplitter in the receiver module is used to spectrally separate the backscatter signals at 1064 and 532 nm, and polarizing beamsplitter cubes (PBC) are used to separate the parallel- (P‖) and cross-polarized (P⊥) signals.
  • The residual cross-talk between P‖ and P⊥ due to Rp and Ts are here reduced by means of additional PBCs in each channel behind the first PBC.
  • The reduction in transmittance and reflectance is included in the corresponding sensitivity factors η.

3.4. BERTHA

  • BERTHA was not designed to provide high-quality polarization measurements.
  • The goal of BERTHA’s depolarization measurements at 710 nm is to separate water from ice clouds, to identify mixed-phase clouds and to discriminate Tellus 61B (2009), 1 desert dust from other less depolarizing aerosols like urban haze or maritime air.
  • After collection with the primary telescope mirror, the photons have to pass five optical elements (reflecting mirrors and dichroic beamsplitters) before reaching the PBC, which reflects the parallel- (P‖) and transmits the cross-polarized signal components at ∼710 nm (eq. 15).
  • The 45◦-calibrations with a polarizing sheet filter before the PBC were performed during several SAMUM evening lidar sessions.

3.5. Sunphotometer and radiosonde

  • Two sunphotometers were installed close to the ground based lidar systems (SSARA and AERONET; see Toledano et al., 2008, and D. Müller, personal communication, 2008, respectively).
  • SSARA measurements are available in one-minute steps and were therefore used for comparison with the MULIS measurements.
  • The AERONET data was used for the evaluation of the DLR-HSRL 1064 nm channel .
  • The Ångström exponent (AE) is derived from the wavelength- (λ) dependent aerosol (i.e. particle) optical depth (AOD) from the fomula AOD (λ) = const.

4. Observations

  • The paper presents linear particle depolarization ratio δp measurements at four wavelengths.
  • The lidar ratios S used for the depolarization retrieval of MULIS at 532 nm were adopted from the coincident DLR-HSRL measurements (displayed in Fig. 6 as broken lines) with errors in the range of ±5 sr.
  • For the whole wavelength range between 355 and 1064 nm the minimum and maximum δp values in Fig. 7 (including the error bars) are confined between 0.17 and 0.39.
  • The AOD and AE values and AE errors in Table 2 are the mean values and standard deviations, respectively, over the periode of the MULIS measurement, as indicated in Table 2.

5. Summary

  • The authors report measurements of the linear particle depolarization ratios δp of pure Saharan dust with four lidar systems at four wavelengths in Quarzazate, Morocco, during SAMUM in May–June 2006.
  • The authors evaluate the errors and their sources of the calibration methods used for the different lidars, and achieve trustworthy error estimations for the linear depolarization ratios.
  • The confidence in the procedures is confirmed by the agreement of the δp values of MULIS and DLR-HSRL within the error bars at 532 nm.
  • The uncertainty of δp comes primarily, at least in their data set, from systematic errors in the setup of the lidar systems, which cannot be reduced by statistical methods.
  • The comparison of the whole δp data set (532 nm) with the corresponding AEs (440–870 nm) from the sunphotometer SSARA exhibits a negative correlation, with AEs between 0.04 and 0.34 for the pure dust cases and between 0.65 and 1.00 for the intermediate periods.

6. Acknowledgments

  • The authors are grateful to the Moroccan Ministry for Foreign Affairs and the Ministry of the Interior for the permission to carry out the SAMUM field campaign in Morocco.
  • The authors would like to extend their gratefulness to the Moroccan Airport Authority and, in particular, to respectable Monsieur Mohammed El Mardi, commander of Ouarzazate airport, for their extraordinary and unbureaucratic support of the participants of SAMUM.
  • The authors would also like to thank Nabil Bousselham for his dedicated support of the research teams.
  • The SAMUM research group is funded by the Deutsche Forschungsgemeinschaft (DFG) under grant number FOR 539.
  • The authors further thank the Johannes Gutenberg University Mainz for its financial support through the research funds of the University of Mainz.

Did you find this useful? Give us your feedback

Figures (12)

Content maybe subject to copyright    Report

Tellus (2009), 61B, 165–179
C
2008 The Authors
Journal compilation
C
2008 Blackwell Munksgaard
Printed in Singapore. All rights reserved
TELLUS
Depolarization ratio profiling at several wavelengths
in pure Saharan dust during SAMUM 2006
By VOLKER FREUDENTHALER
1
,MICHAEL ESSELBORN
2
, MATTHIAS WIEGNER
1
,
BIRGIT HEESE
3
,MATTHIAS TESCHE
3
,ALBERT ANSMANN
3
,DETLEF M
¨
ULLER
3
,
DIETRICH ALTHAUSEN
3
,MARTIN WIRTH
2
,ANDREAS FIX
2
,GERHARD EHRET
2
,
PETER KNIPPERTZ
4
,CARLOS TOLEDANO
1
,JOSEF GASTEIGER
1
,
MARKUS GARHAMMER
1
andMEINHARD SEEFELDNER
1
,
1
Meteorological Institute,
Ludwig-Maximilians-Universit
¨
at, Theresienstr. 37, 80333 Munich, Germany;
2
Institute of Atmospheric Physics,
German Aerospace Center (DLR), Oberpfaffenhofen, 82234 Wessling, Germany;
3
Leibniz Institute for Tropospheric
Research, Permoserstr. 15, 04318 Leipzig, Germany;
4
Institute for Atmospheric Physics, Johannes Gutenberg
University, Becherweg 21, 55099 Mainz, Germany
(Manuscript received 2 January 2008; in final form 18 August 2008)
ABSTRACT
Vertical profiles of the linear particle depolarization ratio of pure dust clouds were measured during the Saharan Mineral
Dust Experiment (SAMUM) at Ouarzazate, Morocco (30.9
N, –6.9
E), close to source regions in May–June 2006,
with four lidar systems at four wavelengths (355, 532, 710 and 1064 nm). The intercomparison of the lidar systems is
accompanied by a discussion of the different calibration methods, including a new, advanced method, and a detailed
error analysis. Over the whole SAMUM periode pure dust layers show a mean linear particle depolarization ratio at
532 nm of 0.31, in the range between 0.27 and 0.35, with a mean Ångstr
¨
om exponent (AE, 440–870 nm) of 0.18 (range
0.04–0.34) and still high mean linear particle depolarization ratio between 0.21 and 0.25 during periods with aerosol
optical thickness less than 0.1, with a mean AE of 0.76 (range 0.65–1.00), which represents a negative correlation of
the linear particle depolarization ratio with the AE. A slight decrease of the linear particle depolarization ratio with
wavelength was found between 532 and 1064 nm from 0.31 ± 0.03 to 0.27 ± 0.04.
1. Introduction
Shape, size distribution and composition of aerosol particles
influence their scattering characteristics and thus the radiative
impact. The polarization lidar technique (Schotland et al., 1971;
Sassen, 1991; Sassen, 2005) is a well-established method to
distinguish ice clouds from water clouds and to identify layers
with ice crystals in mixed–phase clouds (e.g. Ansmann et al.,
2005, 2007). Freudenthaler et al. (1996) applied a scanning po-
larization lidar to study the evolution of contrails. The technique
has been used to identify the type of polar stratospheric clouds
(Reichardt et al., 2000; Toon et al., 2000; Sassen, 2005) and
volcanic ash in the troposphere and stratosphere (Hayashida
et al., 1984; Winker and Osborn, 1992; Sassen et al., 2007).
The polarization lidar is also well suited for aerosol profiling
(McNeil and Carswell, 1975; Murayama et al., 1996; Gobbi,
Corresponding author.
e-mail: volker.freudenthaler@meteo.physik.uni-muenchen.de
DOI: 10.1111/j.1600-0889.2008.00396.x
1998; Cairo et al., 1999; Sassen et al., 2007) and allows us
to unambiguously discriminate desert dust from other aerosols
(Gobbi et al., 2000; di Sarra et al., 2001; Sakai et al., 2002; M
¨
uller
et al., 2003; Iwasaka et al., 2003; Murayama et al., 2004). Based
on model calculations, it has been demonstrated that the spectral
dependence of the dust linear depolarization ratio is sensitive to
the size distribution of the nonspherical scatterers (Mishchenko
and Sassen, 1998). Thus, observations of the linear depolar-
ization ratio at several wavelengths may be used in retrieval
schemes (Dubovik et al., 2006) to improve the estimation of
the microphysical properties of dust from optical measurements
(D. M
¨
uller, personal communication, 2008; Wiegner et al.,
2008). First dual-wavelength aerosol polarization lidar measure-
ments were presented by Sugimoto et al. (2002).
The linear depolarization ratio δ is defined as the ratio of
the cross–polarized lidar return signal to the parallel-polarized
backscatter signal. The planes of polarization of the employed
two detectors are parallel and orthogonal to the plane of polar-
ization of the transmitted linearly polarized laser. The method is
discussed in detail in Section 2. Although the polarization lidar
Tellus 61B (2009), 1 165
PUBLISHED BY THE INTERNATIONAL METEOROLOGICAL INSTITUTE IN STOCKHOLM
SERIES B
CHEMICAL
AND PHYSICAL
METEOROLOGY

166 V. FREUDENTHALER ET AL.
technique appears to be rather simple, and thus robust and reli-
able, several sources for systematic errors can have a significant
impact on the result and can introduce large errors. These error
sources are discussed in Section 3. On the other hand, rather ac-
curate values of the depolarization ratio of desert dust (including
the spectral slope of the ratio) are required for further use in the
above mentioned retrieval schemes. An extensive error analysis
is thus as important as the measurement itself. Several papers
(Biele et al., 2000; Reichardt et al., 2003; Alvarez et al., 2006)
show how highly accurate depolarization observations can be
realized. In this paper, a further technique is discussed, which
improves the calibration of the depolarization channels.
We begin with the presentation of the polarization lidar tech-
nique (methodology, lidar setup) in Sections 2 and 3. In Section
4, the observations are discussed. For the first time, dust depo-
larization ratios at four wavelengths were measured. The unique
field site was close to the source of the dust particles so that
the depolarization properties of pure dust could be investigated.
Section 5 summarizes the main findings, and the error analysis
is treated in the Appendix.
2. Polarization lidar method
Measurements of the linear depolarization ratio δ with lidars are
often performed with the aim to just discern between the dry,
the liquid and the ice phase of aerosols and clouds in the profiles
of one lidar system, which requires only a relative measure of
δ with a low accuracy of the absolute values. During Saharan
Mineral Dust Experiment (SAMUM), it was attempted to mea-
sure a possible wavelength dependence of the dust particle linear
depolarization ratio δ
p
, with four different lidar systems at four
wavelengths as inputs for model calculations of δ
p
regarding the
particles shapes and size distribution. Thus, the uncertainty of
the absolute values must be known and should be small com-
pared with the expected natural variance. The total backscattered
power P(r) with their dependence on the distance r from the lidar
is described by the lidar equation
P =
ηβ (r)τ
2
(r)
r
2
, (1)
where η is the system constant, β the backscatter coefficient,
and the factor τ
2
accounts for the atmospheric transmittance
on the way from the lidar to the scattering volume, and back.
For the determination of δ the lidars used in this study measure
the atmospheric backscatter signals in two receiver channels,
parallel- and cross-polarized with respect to the plane of the
linear polarized output of the laser beam. The two polarization
components are separated in the receiver by means of polarizing
beamsplitter cubes (PBC). But this separation is not perfect. Fur-
thermore the polarizing beamsplitter might be misaligned with
respect to the plane of polarization of the emitted laser beam,
and additionally, a rotation of the polarization plane is used for
the relative calibration of the two receiver channels. Therefore,
Fig. 1. Signal power components in a receiver of a depolarization lidar
with a polarizing beamsplitter cube with reflectivities R
p
and R
s
and
transmittances T
p
and T
s
for linearly polarized light parallel (p) and
perpendicular (s) to the incident plane of the polarizing beamsplitter.
P
R
and P
T
are the measured quantities in the reflected and transmitted
path, respectively, and V
R
and V
T
are the corresponding amplification
factors including the optical transmittances.
we show the necessary equations of the angle ϕ between the
plane of polarization of the laser and the incident plane of the
polarizing beamsplitter cube, according to Fig. 1.
The backscatter powers before the PBC are (skipping the
range dependence in the following for convenience)
P
=
η
β
p
+ β
m
τ
2
r
2
,
P
=
η
β
p
+ β
m
τ
2
r
2
,
(2)
with the system constants η
||
and η
including here only the laser
power and the telescope aperture, assuming negligible diattenu-
ation of the optics before the PBC, for example, a telescope or
dichroic beamsplitters. The backscatter coefficient β is split up
in the parallel- (β
) and cross-polarized (β
) components of the
backscatter from particles (β
p
) and from molecules (β
m
). The
total backscatter power P and the total backscatter coefficient β
are the sum of both polarized components:
P = P
+ P
. (3)
The ratio of the total backscatter coefficient to the molecular
component is called the backscatter ratio R
R =
β
m
+ β
p
β
m
, (4)
and the ratio of the total cross- to the total parallel-polarized
backscatter coefficient is called the linear volume depolarization
ratio δ
v
:
δ
v
=
β
β
=
P
P
. (5)
Tellus 61B (2009), 1

DEPOLARIZATION RATIO PROFILING IN PURE SAHARAN DUST 167
The power components with respect to the incident plane of
the PBC are
P
s
(ϕ) = P
sin
2
(ϕ) + P
cos
2
(ϕ),
P
p
(ϕ) = P
cos
2
(ϕ) + P
sin
2
(ϕ).
(6)
The subscripts p and s denote the planes parallel and perpen-
dicular to the incident plane of the PBC (see Fig. 1), respectively,
and ϕ is the angle between the plane of polarization of the laser
and the incident plane of the PBC. Depending on this angle, the
cross polarized signal P
can be measured in the reflected (for
ϕ = 0
) or in the transmitted path (ϕ = 90
). Hence, we denote
the power measured in the reflected and transmitted paths with
the subscripts R and T, respectively. Behind the PBC the total
reflected (P
R
) and transmitted (P
T
) power components are
P
R
(ϕ) = [P
p
(ϕ)R
p
+ P
s
(ϕ)R
s
]V
R
,
P
T
(ϕ) = [P
p
(ϕ)T
p
+ P
s
(ϕ)T
s
]V
T
.
(7)
The amplification factors V
R
and V
T
include the optical trans-
mittances of the receiver and the electronic amplification in each
channel. P
R
and P
T
are the quantities we actually record with the
data acquisition. For the following it is convenient to introduce a
relative amplification factor V
and the measured signal ratio δ
δ
(ϕ) =
P
R
(ϕ)
P
T
(ϕ)
,V
=
V
R
V
T
. (8)
With eqs. (6)–(8), we achieve
δ
(ϕ) = V
[1 + δ
v
tan
2
(ϕ)]R
p
+ [tan
2
(ϕ) + δ
v
]R
s
[1 + δ
v
tan
2
(ϕ)]T
p
+ [tan
2
(ϕ) + δ
v
]T
s
. (9)
2.1. Calibration
To retrieve the total backscatter power P and δ
v
from the mea-
surements P
R
and P
T
with eqs. (3)–(7), we need V
, which we
can get from calibration measurements in different ways and
using eq. (9) in the following form
V
=
[1 + δ
v
tan
2
(ϕ)]T
p
+ [tan
2
(ϕ) + δ
v
]T
s
[1 + δ
v
tan
2
(ϕ)]R
p
+ [tan
2
(ϕ) + δ
v
]R
s
δ
(ϕ). (10)
For ϕ = 0
, it follows from eq. (10)
V
=
T
p
+ δ
v
T
s
R
p
+ δ
v
R
s
δ
(
0
)
. (11)
With known δ
v
in some range of the lidar signal, we can
determine V
already from a regular measurement with eq.
(11), which we call the ‘0
-calibration’. The linear depolar-
ization ratio δ
m
of air molecules is well known (Behrendt and
Nakamura, 2002), and thus, we could use an aerosol-free li-
dar range in the free troposphere were δ
v
= δ
m
. But already
a very low amount of strong depolarizing aerosols, like dust
or ice crystals in the assumed clean range, causes large er-
rors of δ
v
and thus of V
. Furthermore, several instrumental
Fig. 2. Relative errors of V
over the calibration angle error γ (see
text) calculated using eqs. (9), (10), (12) and (13), with T
p
=0.95, R
p
=
0.05, R
s
= 0.99 and T
s
= 0.01 for δ
v
= 0.0036 (clean air) and δ
v
=
0.30 (desert dust). The ±45
-calibration errors are multiplied by a
factor of 100.
uncertainties can add large errors, especially with this cali-
bration method as described in the Appendix. More reliable
calibration methods use the fact that tan
2
(±45
) = 1, which
makes P
p
= P
s
in eq. (6), and from eq. (9) we get for
ϕ =+45
or ϕ = –45
,
V
=
T
p
+ T
s
R
p
+ R
s
δ
(
±45
)
, (12)
which is independent of δ
v
. We call this method the ‘45
-
calibration’. However, it is difficult to measure the exact angle
of the plane of polarization of a laser beam and the alignment
relative to the polarizing beamsplitter cube. An error γ from
ϕ 45
, of the order of 1
, has to be assumed causing a large
error in V
depending on δ
v
(see Fig. 2). But if we calculate V
from two subsequent measurements at exactly 90
difference,
that is, ϕ =+45
+ γ and ϕ = –45
+ γ :
V
=
T
p
+ T
s
R
p
+ R
s
δ
(
+45
)
× δ
(
45
)
, (13)
it can be shown that the errors compensate each other very well
over a large range of γ (see Fig. 2). We call this method the
±45
-calibration’. The exact 90
difference can be achieved
with high accuracy by means of, for example, mechanical limit
stops (portable lidar system, POLIS; Deutsches Zentrum f
¨
ur
Luft- und Raumfahrt, High Spectral Resolution Lidar, DLR-
HSRL) or a motorized rotation mount (multichannel lidar sys-
tem, MULIS).
The 45
-calibration is based on the fact that P
p
= P
s
are made
equal, which can also be achieved by using a polarizing sheet
filter in front of the PBC at ϕ =+45
or 45
(see Fig. 3). If
the polarizing filter has a negligible small transmission in the
cross-polarized plane (k2 < 10
4
), eqs. (12) and (13) are still
valid, and the calibration errors are similar to those in Fig. 2.
Tellus 61B (2009), 1

168 V. FREUDENTHALER ET AL.
Fig. 3. Signal power components in a receiver of a depolarization lidar
with an additional linear analyser (polarizing sheet filter) with
transmittances parallel (k1) and perpendicular (k2) to the incident
linear polarized light. The other components are as described in Fig. 1.
MULIS and POLIS used the advanced two-angle ±45
-
calibration method with different techniques for the rotation
of the polarization, and the DLR-HSRL used the one angle
45
-calibration method. For Backscatter Extinction lidar–Ratio
Temperature Humidity profiling Apparatus (BERTHA), the 45
-
calibration with a polarizing sheet filter was applied.
2.2. Retrieval of the linear volume depolarization
ratio δ
v
Once V
is known, we get δ
v
with eqs. (5) and (6), for a regular
measurements at ϕ = 0
:
δ
v
=
P
P
=
P
s
P
p
= 0
. (14)
As for commercial PBCs, R
s
is usually much closer to 1 than
T
p
, the noise and error caused by the cross-talk from the strong
parallel-polarized signal to the weaker cross-polarized signal are
reduced if the parallel polarized signal is detected in the reflected
s-branch of the PBC. For this setup ϕ = 90
, and we get
δ
v
=
P
P
=
P
p
P
s
= 90
. (15)
From eqs. (5)–(8) follows
P
s
P
p
=
δ
V
T
p
R
p
R
s
δ
V
T
s
(16)
and
P = P
p
+ P
s
=
V
R
s
R
p
P
T
+
T
p
T
s
P
R
V
R
T
p
R
s
R
p
T
s
. (17)
The knowledge of V
R
is not necessary, as we only need a
relative signal for the lidar signal inversion with the Fernald/Klett
retrieval (Klett, 1985; Fernald, 1984), and thus we can set it to
V
R
= 1. In case the parameters of the polarizing beamsplitter
cube are
T
s
= 1 R
s
,R
p
= 1 T
p
, (18)
which can be assumed for commercial PBCs, the total signal P
is retrieved from
P = P
p
+ P
s
= V
P
T
+ P
R
. (19)
2.3. Retrieval of the linear particle depolarization
ratio δ
p
The δ
p
can be calculated from eqs. (2)–(5) using
δ
p
=
β
p
β
p
=
(
1 + δ
m
)
δ
v
R (1 + δ
v
)δ
m
(1 + δ
m
)R (1 + δ
v
)
(20)
(Biele et al., 2000), with the height in dependent linear depolar-
ization ratio of air molecules:
δ
m
=
β
m
β
m
, (21)
which can be determined with high accuracy (Behrendt and
Nakamura, 2002). The backscatter ratio R can be retrieved from
the total signal P using, for example, the Fernald/Klett inversion
with a reference value β
p
(r
0
) at a reference range r
0
and known
range-dependent lidar ratios S
S =
β
P
α
p
, (22)
where α
p
is the particle extinction coefficient. S(r) must
be retrieved by an additional measurement, for example,
with a Raman channel (Tesche et al., 2008) or a HSRL
(Esselborn et al., 2008). The values of δ
v
and R are subject to
systematic and statistical (noise) errors. A detailed error propa-
gation analysis can be found in the Appendix.
3. Polarization lidar in SAMUM
The linear depolarization ratio of the dust particles was mea-
sured during SAMUM with four lidar systems at four wave-
lengths:MULIS at 532 nm;POLIS at 355 nm (both from
Munich);BERTHA at 710 nm (Leipzig) and the airborne DLR
HSRL at 532 and 1064 nm. MULIS and BERTHA were located
at the airport of Quarzazate (1133 m a.s.l., 30.938
N, –6.907
E)
about 10 m apart from each other and POLIS at about 100 m
distance. All three ground based systems can change the zenith
angle of the sounding (scanning capability), which possibility
was used frequently. We assume that the profiles of the different
lidar systems are well comparable as orographical effects of the
surrounding are negligible and the dust layer was mostly suf-
ficiently homogeneous over the considered periods. The DLR-
HSRL was installed on board the DLR Falcon airplane, which
flew several times over the ground based lidars, with high spa-
tial accuracy. For the temporal averaging of these measurements,
care was taken to consider only sections of the time-series with
comparably small changes in the lidar profiles. Owing to a lack of
adequate reference sources, MULIS and POLIS are described in
Tellus 61B (2009), 1

DEPOLARIZATION RATIO PROFILING IN PURE SAHARAN DUST 169
Table 1. System parameters of the depolarization channels of the lidar systems
POLIS MULIS BERTHA DLR-HSRL
Laser/Receiver Biaxial Biaxial Coaxial Coaxial
Laser Brilliant Ultra Continuum, Surelite II Ti:Saphire, Solar TII Ltd. CF 125 Nd:YAG
Wavelengths (nm) 355 532 710 1064, 532
Pulse energy (mJ) 7.8 50 20 220, 100
Repetition rate (Hz) 20 10 30 100
Beam divergence (mrad) 0.69 0.6 <0.3 0.5
Telescope Dall-Kirkham Cassegrain Cassegrain Cassegrain
Diameter (m) 0.2 0.3 0.53 0.35
Focal length (m) 1.2 0.96 3 5
Field stop (mrad) 2.5 0 to 3, adjustable 0.6 1
Receiver optics
Interference filter bandwidth (nm) 1.0, fwhm 1.1, fwhm 0.4, fwhm 1.0, fwhm
PBC T
p
0.9521 0.9831 0.95 1
PBC R
s
0.9985 0.9965 0.999 1
Depol.-calibration method ±45
manual ±45
wave plate 45
polarizer 45
manual
more detail in the following. For the DLR-HSRL and BERTHA,
the details regarding the depolarization measurements comple-
ment the descriptions in the references mentioned below. Table 1
gives a summary of the relevant lidar system parameters.
3.1. MULIS
MULIS is a mobile, five wavelength lidar system with backscat-
ter channels at 355, 532 and 1064 nm and Raman channels at
387 and 607 nm. The Nd:YAG laser fundamental wavelength is
frequency doubled and tripled by means of Potassium Dideu-
terium Phosphate (KD
P) crystals, and the output at 532 nm is
assumed to be perfectly linear polarized. The adjustable filed
stop (TS in Fig. 4) is used to decrease the field of view and with
that, the signal power in the near range during the calibration
measurements, because at ϕ 45
(Fig. 1), the signal in the
cross-polarized channel can increase by an order of magnitude.
Such a high signal exceeds the linear range of the data acquisi-
tion, especially that of the preamplifier, which would introduce
signal distortions and errors in V
, which are very difficult to
assess. To minimize diattenuation and polarization effects of the
dichroic beamsplitters, all optical coatings are custom-designed
(Laseroptik, Germany). For the same reason, the beamsplitters
BS1toBS4areusedat30
reflection angle (see Fig. 4). To re-
duce cross-talk, the custom-made PBC (Optarius, UK) has high
transmittance T
p
and high reflection R
s
,andP
is detected in the
reflected branch P
R
of the PBC (see eq. 15). A cemented, true
zero-order half-wave plate at 532 nm (Casix, China) is used in
front of the PBC to rotate the plane of polarization with a verified
accuracy and repeatability of better than 0.1
. The accuracy of
the ϕ = 0
position is better than ±0.2
,anda±45
-calibration
was performed every time the lidar setup had been changed. The
detectors are Hamamatsu R7400 PMTs in the depolarization
Fig. 4. Optical setup of the MULIS receiver with analog backscatter
channels at 355 nm, 532 nm parallel- and cross- polarized, 1064 nm
and Raman channels at 387 and 607 nm. TS, tilted slit diaphragm; L1,
collimating lens; BS#, dichroic beamsplitters (BS3, 6 and 7 are used as
mirrors); IFF#, interference filter; PBC, polarizing beamsplitter cube at
532 nm and HWP, half-wave plate at 532 nm for the calibration of the
depolarization channels.
channels. The sensitivity of individual channels was adjusted
with absorbing neutral density filters (Schott). The preampli-
fier stages of the 12 bit (P
R
, P
) and 14 bit (P
T
) ADC-boards
(Spectrum GmbH, Germany) were optimized regarding signal
Tellus 61B (2009), 1

Citations
More filters
Journal ArticleDOI
TL;DR: The European Aerosol Research Lidar Network (EARLINET) as mentioned in this paper was founded as a research project for establishing a quantitative, comprehensive, and statistically significant database for the horizontal, vertical, and tempo-ral distribution of aerosols on a continental scale.
Abstract: The European Aerosol Research Lidar Network, EARLINET, was founded in 2000 as a research project for establishing a quantitative, comprehensive, and statistically significant database for the horizontal, vertical, and tempo- ral distribution of aerosols on a continental scale. Since then EARLINET has continued to provide the most extensive col- lection of ground-based data for the aerosol vertical distribu- tion over Europe. This paper gives an overview of the network's main de- velopments since 2000 and introduces the dedicated EAR- LINET special issue, which reports on the present innova- tive and comprehensive technical solutions and scientific re- sults related to the use of advanced lidar remote sensing tech- niques for the study of aerosol properties as developed within the network in the last 13 years. Since 2000, EARLINET has developed greatly in terms of number of stations and spatial distribution: from 17 sta- tions in 10 countries in 2000 to 27 stations in 16 countries in 2013. EARLINET has developed greatly also in terms of technological advances with the spread of advanced multi- wavelength Raman lidar stations in Europe. The develop- ments for the quality assurance strategy, the optimization of instruments and data processing, and the dissemination of data have contributed to a significant improvement of the net- work towards a more sustainable observing system, with an increase in the observing capability and a reduction of oper- ational costs. Consequently, EARLINET data have already been ex- tensively used for many climatological studies, long-range transport events, Saharan dust outbreaks, plumes from vol- canic eruptions, and for model evaluation and satellite data validation and integration. Future plans are aimed at continuous measurements and near-real-time data delivery in close cooperation with other ground-based networks, such as in the ACTRIS (Aerosols, Clouds, and Trace gases Research InfraStructure Network) www.actris.net, and with the modeling and satellite commu- nity, linking the research community with the operational world, with the aim of establishing of the atmospheric part of the European component of the integrated global observ- ing system.

417 citations

Journal ArticleDOI
TL;DR: In this article, a methodology based on observations of known aerosol types is used to qualitatively classify the extensive set of airborne high-spectral resolution lidar measurements into eight separate types.
Abstract: . The NASA Langley Research Center (LaRC) airborne High Spectral Resolution Lidar (HSRL) on the NASA B200 aircraft has acquired extensive datasets of aerosol extinction (532 nm), aerosol optical depth (AOD) (532 nm), backscatter (532 and 1064 nm), and depolarization (532 and 1064 nm) profiles during 18 field missions that have been conducted over North America since 2006. The lidar measurements of aerosol intensive parameters (lidar ratio, depolarization, backscatter color ratio, and spectral depolarization ratio) are shown to vary with location and aerosol type. A methodology based on observations of known aerosol types is used to qualitatively classify the extensive set of HSRL aerosol measurements into eight separate types. Several examples are presented showing how the aerosol intensive parameters vary with aerosol type and how these aerosols are classified according to this new methodology. The HSRL-based classification reveals vertical variability of aerosol types during the NASA ARCTAS field experiment conducted over Alaska and northwest Canada during 2008. In two examples derived from flights conducted during ARCTAS, the HSRL classification of biomass burning smoke is shown to be consistent with aerosol types derived from coincident airborne in situ measurements of particle size and composition. The HSRL retrievals of AOD and inferences of aerosol types are used to apportion AOD to aerosol type; results of this analysis are shown for several experiments.

393 citations

Journal ArticleDOI
TL;DR: In this paper, a case study of the Saharan Mineral Dust Experiment (SAMUM) was presented to discuss the potential of the technique and compared with column-integrated products obtained with Aerosol Robotic Network Sun photometer.
Abstract: [1] Multiwavelength aerosol Raman lidar in combination with polarization lidar at Praia (14.9N, 23.5W), Cape Verde, is used to separate the optical properties of desert dust and biomass burning particles as a function of height in the mixed dust and smoke plumes over the tropical North Atlantic west of the African continent. The advanced lidar method furthermore permits the derivation of the single-scattering albedo and microphysical properties of the African biomass burning smoke. A case study is presented to discuss the potential of the technique. The observations were performed during the Saharan Mineral Dust Experiment (SAMUM) in January and February 2008. The height-resolved lidar results are compared with column-integrated products obtained with Aerosol Robotic Network Sun photometer. Good agreement is found. Furthermore, the findings are compared with lidar and aircraft observations recently performed in western Africa and with our previous lidar observations taken in tropical and subtropical regions of southern and eastern Asia. The SAMUM case study represents typical aerosol layering conditions in the tropical outflow regime of western Africa during winter season. Above a dense desert dust layer (with an optical depth of about 0.25 at 532 nm) which reached to 1500 m, a lofted layer consisting of desert dust (0.08 optical depth) and biomass burning smoke (0.24 optical depth) extended from 1500 to 5000 m height. Extinction values were 20 ± 10 Mm � 1 (desert dust) and 20–80 Mm � 1 (smoke) in the lofted plume. The smoke extinction-to-backscatter ratios were rather high, with values up to more than 100 sr, effective radii ranged from 0.15 to 0.35 mm, and the smoke single-scattering albedo was partly below 0.7.

294 citations

Journal ArticleDOI
01 Sep 2011-Tellus B
TL;DR: In this paper, the particle linear depolarization ratio of Saharan dust, marine aerosols and mixtures of biomass-burning aerosols from southern West Africa and Saharan Dust was determined at three wavelengths with three lidar systems during the SAharan Mineral dUst experiMent 2 at the airport of Praia, Cape Verde, between 22 January and 9 February 2008.
Abstract: The particle linear depolarization ratio δ p of Saharan dust, marine aerosols and mixtures of biomass-burning aerosols from southern West Africa and Saharan dust was determined at three wavelengths with three lidar systems during the SAharan Mineral dUst experiMent 2 at the airport of Praia, Cape Verde, between 22 January and 9 February 2008. The lidar ratio S p of these major types of tropospheric aerosols was analysed at two wavelengths. For Saharan dust, we find wavelength dependent mean particle linear depolarization ratios δ p of 0.24–0.27 at 355 nm, 0.29–0.31 at 532 nm and 0.36–0.40 at 710 nm, and wavelength independent mean lidar ratios S p of 48–70 sr. Mixtures of biomass-burning aerosols and dust show wavelength independent values of δ p and S p between 0.12–0.23 and 57–98 sr, respectively. The mean values of marine aerosols range independent of wavelength for δ p from 0.01 to 0.03 and for S p from 14 to 24 sr. DOI: 10.1111/j.1600-0889.2011.00556.x

239 citations


Cites background or methods from "Depolarization ratio profiling at s..."

  • ...Measurements at four wavelengths of fresh Saharan dust during SAMUM 1 were presented in Freudenthaler et al. (2009) with mean values of 0.27 < δp < 0.35 at 532 nm....

    [...]

  • ...The depolarization calibration of all three lidar systems is described in Freudenthaler et al. (2009)....

    [...]

  • ...The error analysis is done according to Freudenthaler et al. (2009)....

    [...]

  • ..., 2008) at 355 nm, the MUlti wavelength LIdar System (MULIS; Freudenthaler et al., 2009) at 532 nm, both from the Meteorological Institute of the Ludwig-Maximilians-Universität, München (MIM), and the multi-wavelength system BERTHA (Backscatter Extinction lidar Ratio Temperature Humidity profiling Apparatus; Althausen et al....

    [...]

  • ...Measurements of δp of mineral dust were performed at four wavelengths over Ouarzazate, Morocco (Freudenthaler et al., 2009), during the first campaign of the SAharan Mineral dUst experiMent (SAMUM; Heintzenberg, 2009)....

    [...]

Journal ArticleDOI
01 Sep 2011-Tellus B
TL;DR: The Saharan Mineral Dust Experiment (SAMUM) project as discussed by the authors has been used to study the impact of Saharan dust on radiative transfer and the feedback of radiative effects upon dust emission and aerosol transport.
Abstract: Two comprehensive field campaigns were conducted in 2006 and 2008 in the framework of the Saharan Mineral Dust Experiment (SAMUM) project. The relationship between chemical composition, shape morphology, size distribution and optical effects of the dust particles was investigated. The impact of Saharan dust on radiative transfer and the feedback of radiative effects upon dust emission and aerosol transport were studied. Field observations (ground-based, airborne and remote sensing) and modelling results were compared within a variety of dust closure experiments with a strong focus on vertical profiling. For the first time, multiwavelength Raman/polarization lidars and an airborne high spectral resolution lidar were involved in major dust field campaigns and provided profiles of the volume extinction coefficient of the particles at ambient conditions (for the full dust size distribution), of particle-shape-sensitive optical properties at several wavelengths, and a clear separation of dust and smoke profiles allowing for an estimation of the single-scattering albedo of the biomass-burning aerosol. SAMUM–1 took place in southern Morocco close to the Saharan desert in the summer of 2006, whereas SAMUM–2 was conducted in Cape Verde in the outflow region of desert dust and biomass-burning smoke from western Africa in the winter of 2008. This paper gives an overview of the SAMUM concept, strategy and goals, provides snapshots (highlights) of SAMUM–2 observations and modelling efforts, summarizes main findings of SAMUM–1 and SAMUM–2 and finally presents a list of remaining problems and unsolved questions.

233 citations


Cites background from "Depolarization ratio profiling at s..."

  • ...Field campaigns are required to validate such models (Heinold et al., 2009; Müller et al., 2009a; Johnson et al., 2011) as well as satellite remote sensing retrievals (Kahn et al., 2009; Dinter et al., 2009; Christopher et al., 2008), including the CALIPSO lidar retrievals (Wandinger et al., 2010)....

    [...]

References
More filters
Journal ArticleDOI
TL;DR: A restatement of the more general solution of Fernald et al.l which is also applicable to mildly turbid atmospheres where both aerosol and molecular scatterers must be considered in the analysis.
Abstract: There have been many discussions of solutions to the lidar equation for elastic scattering (e.g., Fernald et al.,' Klett, 2 Davis, and Collis and Russell ). Most of these are simply variations on Hitschfeld and Bordan's5 solution for meteorological radars. Klett 2 recently restated this solution in a very convenient form for the analysis of lidar observations collected in very turbid atmospheres. His paper has prompted a restatement of the more general solution of Fernald et al.l which is also applicable to mildly turbid atmospheres where both aerosol and molecular scatterers must be considered in the analysis. This has led to a simple numerical scheme for the computer analysis of lidar measurements. The lidar equation for two distinct classes of scatters (Fernald et al.') is

1,558 citations


"Depolarization ratio profiling at s..." refers methods in this paper

  • ...(17) The knowledge of VR is not necessary, as we only need a relative signal for the lidar signal inversion with the Fernald/Klett retrieval (Klett, 1985; Fernald, 1984), and thus we can set it to VR = 1....

    [...]

  • ...The knowledge of VR is not necessary, as we only need a relative signal for the lidar signal inversion with the Fernald/Klett retrieval (Klett, 1985; Fernald, 1984), and thus we can set it to VR = 1....

    [...]

Journal ArticleDOI
TL;DR: In this article, the authors used shape mixtures of randomly oriented spheroids for modeling desert dust aerosol light scattering, and the results indicated that nonspherical particles with aspect ratios similar to 1.5 dominate in desert dust plumes, while in the case of background maritime aerosol spherical particles are dominant.
Abstract: [ 1] The possibility of using shape mixtures of randomly oriented spheroids for modeling desert dust aerosol light scattering is discussed. For reducing calculation time, look-up tables were simulated for quadrature coefficients employed in the numerical integration of spheroid optical properties over size and shape. The calculations were done for 25 bins of the spheroid axis ratio ranging from similar to 0.3 ( flattened spheroids) to similar to 3.0 ( elongated spheroids) and for 41 narrow size bins covering the size parameter range from similar to 0.012 to similar to 625. The look-up tables were arranged into a software package, which allows fast, accurate, and flexible modeling of scattering by randomly oriented spheroids with different size and shape distributions. In order to evaluate spheroid model and explore the possibility of aerosol shape identification, the software tool has been integrated into inversion algorithms for retrieving detailed aerosol properties from laboratory or remote sensing polarimetric measurements of light scattering. The application of this retrieval technique to laboratory measurements by Volten et al. ( 2001) has shown that spheroids can closely reproduce mineral dust light scattering matrices. The spheroid model was utilized for retrievals of aerosol properties from atmospheric radiation measured by AERONET ground-based Sun/sky-radiometers. It is shown that mixtures of spheroids allow rather accurate fitting of measured spectral and angular dependencies of observed intensity and polarization. Moreover, it is shown that for aerosol mixtures with a significant fraction of coarse-mode particles ( radii >= similar to 1 mu m), the nonsphericity of aerosol particles can be detected as part of AERONET retrievals. The retrieval results indicate that nonspherical particles with aspect ratios similar to 1.5 and higher dominate in desert dust plumes, while in the case of background maritime aerosol spherical particles are dominant. Finally, the potential of using AERONET derived spheroid mixtures for modeling the effects of aerosol particle nonsphericity in other remote sensing techniques is discussed. For example, the variability of lidar measurements ( extinction to backscattering ratio and signal depolarization ratio) is illustrated and analyzed. Also, some potentially important differences in the sensitivity of angular light scattering to parameters of nonspherical versus spherical aerosols are revealed and discussed.

1,260 citations


"Depolarization ratio profiling at s..." refers background in this paper

  • ...Thus, observations of the linear depolarization ratio at several wavelengths may be used in retrieval schemes (Dubovik et al., 2006) to improve the estimation of the microphysical properties of dust from optical measurements (D....

    [...]

  • ...Thus, observations of the linear depolarization ratio at several wavelengths may be used in retrieval schemes (Dubovik et al., 2006) to improve the estimation of the microphysical properties of dust from optical measurements (D. Müller, personal communication, 2008; Wiegner et al., 2008)....

    [...]

Journal ArticleDOI
TL;DR: An alternative formulation is given herein which assumes the proportionality factor in the power law relationship is itself a function of range or extinction, and the resulting lidar equation is solvable.
Abstract: The conventional approach to solving the single-scattering lidar equation makes use of the assumption of a power law relation between backscatter and extinction with a fixed exponent and constant of proportionality. An alternative formulation is given herein which assumes the proportionality factor in the power law relationship is itself a function of range or extinction. The resulting lidar equation is solvable as before, and examples are given to show how even an approximate description of deviations from the power law form can yield an improved inversion solution for the extinction. A further generalization is given which includes the effects of a background of Rayleigh scatterers.

832 citations


"Depolarization ratio profiling at s..." refers methods in this paper

  • ...(17) The knowledge of VR is not necessary, as we only need a relative signal for the lidar signal inversion with the Fernald/Klett retrieval (Klett, 1985; Fernald, 1984), and thus we can set it to VR = 1....

    [...]

  • ...The knowledge of VR is not necessary, as we only need a relative signal for the lidar signal inversion with the Fernald/Klett retrieval (Klett, 1985; Fernald, 1984), and thus we can set it to VR = 1....

    [...]

Journal ArticleDOI
TL;DR: In this article, the development of the polarization lidar technique is reviewed, and the current capabilities and limitations of the technique for the cloud research are discussed, as well as the current theoretical approaches involving ice crystal ray-tracing and cloud microphysical-model simulations are expected to increase the utility of the Lidar technique.
Abstract: The development of the polarization lidar technique is reviewed, and the current capabilities and limitations of the technique for the cloud research are discussed. At present, polarization lidar is a key component of climate-research programs designed to characterize the properties of cirrus clouds and is an integral part of multiple remote-sensor studies of mixed-phase cloud systems such as winter mountain storms, making it possible to discriminate between cloud phases and to identify some particle types and orientations. Recent theoretical approaches involving ice crystal ray-tracing and cloud microphysical-model simulations are expected to increase the utility of the polarization lidar technique.

524 citations


"Depolarization ratio profiling at s..." refers methods in this paper

  • ...The polarization lidar technique (Schotland et al., 1971; Sassen, 1991; Sassen, 2005) is a well-established method to distinguish ice clouds from water clouds and to identify layers with ice crystals in mixed–phase clouds (e....

    [...]

  • ...The polarization lidar technique (Schotland et al., 1971; Sassen, 1991; Sassen, 2005) is a well-established method to distinguish ice clouds from water clouds and to identify layers with ice crystals in mixed–phase clouds (e.g. Ansmann et al., 2005, 2007)....

    [...]

Journal ArticleDOI
TL;DR: In this paper, the authors used polarization lidars for continuous observations of aerosols in China and Japan during March to May 2001, corresponding with the Asian Pacific Regional Aerosol Characterization Experiment (ACE-Asia) field campaign period.
Abstract: [1] Continuous observations of aerosols in China and Japan were made by polarization lidars during March to May 2001, corresponding with the Asian Pacific Regional Aerosol Characterization Experiment (ACE-Asia) field campaign period. Lidars in Beijing, Nagasaki, and Tsukuba were continuously operated regardless of weather conditions. Scatterers in the atmosphere were categorized for all vertical profiles, and occurrence frequencies of dust, spherical aerosols, and clouds up to 6 km were calculated. The frequency of dust was highest in Beijing for the whole height range. There was a peak of dust occurrence near the ground in Nagasaki. Dust was frequently detected in the free troposphere in Tsukuba. The contributions of dust and spherical aerosols to the total backscattering coefficient were estimated from the depolarization ratio with the assumption of the external mixture of both kinds of aerosols. Vertical profiles of backscattering by dust and by spherical aerosols represented the different characteristics of these aerosols. The monthly averaged backscattering coefficients by dust near the surface were 0.003/km/sr in Beijing, 0.001–0.002/km/sr in Nagasaki, and 0.0006/km/sr in Tsukuba. The backscattering coefficients by spherical aerosols near the surface were 0.002–0.004/km/sr at all three observatories. We compared the derived backscattering coefficients with aerosol mass concentrations calculated by a numerical model, Chemical Weather Forecasting System (CFORS). CFORS reproduced well the vertical structures of the tall dust events and the enhancements of spherical aerosols throughout the observation period. A specific dust event on 16–19 May 2001 was analyzed by using five lidars in Japan, and its fine structure is described.

431 citations


"Depolarization ratio profiling at s..." refers background in this paper

  • ...…partly or completely mixed with maritime particles, anthropogenic pollution or biomass-burning smoke (e.g. Sakai et al., Tellus 61B (2009), 1 2000; Sakai et al., 2002; Shimizu et al., 2004; Sugimoto et al., 2002; Ansmann et al., 2003; Murayama et al., 2004; Müller et al., 2003; Chen et al., 2007)....

    [...]

Frequently Asked Questions (12)
Q1. What contributions have the authors mentioned in the paper "Depolarization ratio profiling at several wavelengths in pure saharan dust during samum 2006" ?

Freudenthaler et al. this paper used a scanning polarization lidar to study the evolution of contrails. 

Using microphysical properties as input parameters, the optical properties of mineral dust, including the linear particle depolarization ratio δp, can be modelled by means of T-matrix method, assuming that the particles can be approximated by spheroids (Wiegner et al., 2008). 

size distribution and composition of aerosol particles influence their scattering characteristics and thus the radiative impact. 

It turns out that the main error source is not the signal noise, as it can, in general, be reduced sufficiently by means of spatial or temporal averaging. 

Due to weak signals in the Raman channels and the limited dynamic range of the photocounting detection, the extinction coefficient α and the lidar ratio S were not retrieved from MULIS measurements but were adopted from the much more accurate DLR-HSRL measurements (Esselborn et al., 2008) or from BERTHA (Tesche et. al., 2008). 

Focusing on Saharan dust, Gobbi et al. (2000) show for backscatter ratios R of 4 and more linear volume depolarization ratios δv of 0.3 during a campaign on Crete, Greece, about 500 km north of Africa. 

The Ångström exponent (AE) is derived from the wavelength- (λ) dependent aerosol (i.e. particle) optical depth (AOD) from the fomulaAOD (λ) = const. 

The ratio of the total backscatter coefficient to the molecular component is called the backscatter ratio RR = β m + βp βm , (4)and the ratio of the total cross- to the total parallel-polarized backscatter coefficient is called the linear volume depolarization ratio δv:δv = β⊥ β‖ = P⊥ P‖ . 

The uncertainty of δp comes primarily, at least in their data set, from systematic errors in the setup of the lidar systems, which cannot be reduced by statistical methods. 

The authors found the main error sources to originate from the depolarization calibration (V∗), with large differences between the different calibration methods, and from the error of the particle backscatter coefficient βp (or backscatter ratio R) due to the uncertainty in the height-dependent lidar ratio S(r) and the uncertainty in the signal calibration in the assumed clean, free troposphere βp(r0). 

The optical elements, located between the P‖- and the Pp-planes before the polarizing sheet filter in Fig. 3, show a significant diattenuation D = 0.726 ± 2%, which is not included in the depolarization calibration but was determined with separate measurements. 

Based on model calculations, it has been demonstrated that the spectral dependence of the dust linear depolarization ratio is sensitive to the size distribution of the nonspherical scatterers (Mishchenko and Sassen, 1998).