COSMIC GPS Radio Occultation Temperature Profiles in Clouds

15
COSMIC GPS Radio Occultation Temperature Profiles in Clouds L. LIN AND X. ZOU The Florida State University, Tallahassee, Florida R. ANTHES University Corporation for Atmospheric Research, Boulder, Colorado Y.-H. KUO National Center for Atmospheric Research, Boulder, Colorado (Manuscript received 3 March 2009, in final form 16 October 2009) ABSTRACT Thermodynamic states in clouds are closely related to physical processes such as phase changes of water and longwave and shortwave radiation. Global Positioning System (GPS) radio occultation (RO) data are not affected by clouds and have high vertical resolution, making them ideally suited to cloud profiling on a global basis. By comparing the Constellation Observing System for Meteorology, Ionosphere, and Climate (COSMIC) RO refractivity data with those of the National Centers for Environmental Prediction–National Center for Atmospheric Research (NCEP–NCAR) reanalysis and ECMWF analysis for soundings in clouds and clear air separately, a systematic bias of opposite sign was found between large-scale global analyses and the GPS RO observations under cloudy and clear-sky conditions. As a modification to the standard GPS RO wet tem- perature retrieval that does not distinguish between cloudy- and clear-sky conditions, a new cloudy retrieval algorithm is proposed to incorporate the knowledge that in-cloud specific humidity (which affects the GPS refractivities) should be close to saturation. To implement this new algorithm, a linear regression model for a sounding-dependent relative humidity parameter a is first developed based on a high correlation between relative humidity and ice water content. In the absence of ice water content information, a takes an empirical value of 85%. The in-cloud temperature profile is then retrieved from GPS RO data modeled by a weighted sum of refractivities with and without the assumption of saturation. Compared to the standard wet retrieval, the cloudy temperature retrieval is consistently warmer within clouds by ;2 K and slightly colder near the cloud top (;1 K) and cloud base (1.5 K), leading to a more rapid increase of the lapse rate with height in the upper half of the cloud, from a nearly constant moist lapse rate below and at the cloud middle (;68C km 21 ) to a value of 7.78C km 21 , which must be closer to the dry lapse rate than the standard wet retrieval. 1. Introduction Clouds play an important role in Earth’s climate, and their modeling, or parameterization, is one of the most uncertain aspects of climate models. They influence the radiative energy budget and atmospheric hydrological processes. Temperatures within clouds are important for determining many other cloud properties, such as cloud water phase, cloud particle size, and cloud water path, whose accurate measurements are required for improve- ments and validations of climate and weather forecast models (Stephens et al. 1990; Diak et al. 1998; Bayler et al. 2000). In the past, a large number of studies have been carried out, including the intercomparisons of sat- ellite cloud-top height retrieval with observations from in situ, airborne, and surface-based instruments (Smith and Platt 1978; Smith and Frey 1990; Frey et al. 1999; Wylie and Wang 1999; Naud et al. 2002, 2004; Hollars et al. 2004; Mahesh et al. 2004; Sherwood et al. 2004; Hawkinson et al. 2005; Stubenrauch et al. 2005; Holz et al. 2006; Kahn et al. 2007; Weisz et al. 2007a,b). The primary data sources for studies related to in-cloud thermodynamic structures are in situ dropsonde measurements from air- craft and weather balloons (radiosondes), and remote Corresponding author address: X. Zou, Department of Meteo- rology, The Florida State University, Tallahassee, FL 32306-4520. E-mail: [email protected] 1104 MONTHLY WEATHER REVIEW VOLUME 138 DOI: 10.1175/2009MWR2986.1 Ó 2010 American Meteorological Society

Transcript of COSMIC GPS Radio Occultation Temperature Profiles in Clouds

Page 1: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

COSMIC GPS Radio Occultation Temperature Profiles in Clouds

L. LIN AND X. ZOU

The Florida State University, Tallahassee, Florida

R. ANTHES

University Corporation for Atmospheric Research, Boulder, Colorado

Y.-H. KUO

National Center for Atmospheric Research, Boulder, Colorado

(Manuscript received 3 March 2009, in final form 16 October 2009)

ABSTRACT

Thermodynamic states in clouds are closely related to physical processes such as phase changes of water

and longwave and shortwave radiation. Global Positioning System (GPS) radio occultation (RO) data are not

affected by clouds and have high vertical resolution, making them ideally suited to cloud profiling on a global

basis. By comparing the Constellation Observing System for Meteorology, Ionosphere, and Climate (COSMIC)

RO refractivity data with those of the National Centers for Environmental Prediction–National Center for

Atmospheric Research (NCEP–NCAR) reanalysis and ECMWF analysis for soundings in clouds and clear air

separately, a systematic bias of opposite sign was found between large-scale global analyses and the GPS RO

observations under cloudy and clear-sky conditions. As a modification to the standard GPS RO wet tem-

perature retrieval that does not distinguish between cloudy- and clear-sky conditions, a new cloudy retrieval

algorithm is proposed to incorporate the knowledge that in-cloud specific humidity (which affects the GPS

refractivities) should be close to saturation. To implement this new algorithm, a linear regression model for

a sounding-dependent relative humidity parameter a is first developed based on a high correlation between

relative humidity and ice water content. In the absence of ice water content information, a takes an empirical

value of 85%. The in-cloud temperature profile is then retrieved from GPS RO data modeled by a weighted

sum of refractivities with and without the assumption of saturation. Compared to the standard wet retrieval,

the cloudy temperature retrieval is consistently warmer within clouds by ;2 K and slightly colder near the

cloud top (;1 K) and cloud base (1.5 K), leading to a more rapid increase of the lapse rate with height in the

upper half of the cloud, from a nearly constant moist lapse rate below and at the cloud middle (;68C km21) to

a value of 7.78C km21, which must be closer to the dry lapse rate than the standard wet retrieval.

1. Introduction

Clouds play an important role in Earth’s climate, and

their modeling, or parameterization, is one of the most

uncertain aspects of climate models. They influence the

radiative energy budget and atmospheric hydrological

processes. Temperatures within clouds are important for

determining many other cloud properties, such as cloud

water phase, cloud particle size, and cloud water path,

whose accurate measurements are required for improve-

ments and validations of climate and weather forecast

models (Stephens et al. 1990; Diak et al. 1998; Bayler

et al. 2000). In the past, a large number of studies have

been carried out, including the intercomparisons of sat-

ellite cloud-top height retrieval with observations from in

situ, airborne, and surface-based instruments (Smith and

Platt 1978; Smith and Frey 1990; Frey et al. 1999; Wylie

and Wang 1999; Naud et al. 2002, 2004; Hollars et al. 2004;

Mahesh et al. 2004; Sherwood et al. 2004; Hawkinson

et al. 2005; Stubenrauch et al. 2005; Holz et al. 2006;

Kahn et al. 2007; Weisz et al. 2007a,b). The primary data

sources for studies related to in-cloud thermodynamic

structures are in situ dropsonde measurements from air-

craft and weather balloons (radiosondes), and remote

Corresponding author address: X. Zou, Department of Meteo-

rology, The Florida State University, Tallahassee, FL 32306-4520.

E-mail: [email protected]

1104 M O N T H L Y W E A T H E R R E V I E W VOLUME 138

DOI: 10.1175/2009MWR2986.1

� 2010 American Meteorological Society

Page 2: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

sensing from weather satellites. However, measurements

from dropsondes and radiosondes have limited Earth

coverage and are not available under severe weather con-

ditions. Operational weather satellites can only identify

horizontal distributions of clouds; they cannot observe the

detailed vertical variability of the thermodynamic struc-

ture within clouds.

The Global Positioning System (GPS) radio occulta-

tion (RO) limb-sounding technique makes use of radio

signals from the GPS satellites for sounding Earth’s at-

mosphere. This technique has the following unique fea-

tures: high accuracy and precision, high vertical resolution,

no contamination from clouds, no system calibration

required, and no instrument drift (Kursinski et al. 1997;

Rocken et al. 1997; Anthes et al. 2000; Kuo et al. 2004).

These features make RO particularly useful for climate

studies. Early studies show that GPS RO soundings from

the first Earth RO mission, GPS Meteorology (GPS/

MET), agree with those from radiosondes and large-scale

analyses to better than 1.5 K between 5- and 30-km alti-

tude (Kursinski et al. 1996; Ware et al. 1996; Rocken et al.

1997; Gorbunov and Kornblueh 2001). Studies by Wickert

et al. (2001, 2004) conclude that the differences between

a later GPS RO mission, Challenging Minisatellite Pay-

load (CHAMP), and the European Centre for Medium-

Range Weather Forecasts (ECMWF) analysis are less

than 1 K above the tropopause and less than 0.5 K in the

altitude range from 12 to 20 km in mid- and high lati-

tudes. Hajj et al. (2004) have compared CHAMP occul-

tation results with another GPS mission, the Argentina’s

Satelite de Aplicaciones Cientificas-C (SAC-C), and con-

cluded that the individual CHAMP profiles are accurate

to better than 0.6 K between 5 and 15 km. Temperature

profiles derived from Constellation Observing System for

Meteorology, Ionosphere, and Climate/Formosa Satellite

Mission 3 (COSMIC/FORMOSAT3, hereafter referred

to as COSMIC for brevity) have even better accuracies

than those of CHAMP. The global mean differences

between COSMIC and high-quality reanalyses within

the height range between 8 and 30 km are estimated

to be ;0.658C (Kishore et al. 2008). The precision of

COSMIC GPS RO soundings, estimated by comparison

of closely collocated COSMIC soundings, is approxi-

mately 0.058C in the upper troposphere and lower

stratosphere (Anthes et al. 2008).

Earlier validations of RO data have been carried out

in terms of refractivity as well as temperature retrieval

and without separating cloudy- and clear-air soundings.

The standard one-dimensional variational data assimi-

lation (1DVar) temperature retrieval did not include

any information on clouds. In this study, we (i) compare

COSMIC RO data (refractivity and temperatures) with

thoseof theNationalCenters forEnvironmentalPrediction–

National Center for Atmospheric Research (NCEP–

NCAR) reanalysis (Kalnay et al. 1996) and ECMWF

analysis from soundings in clouds and clear air sepa-

rately; and (ii) examine whether the standard 1DVar

temperature retrieval results are improved by adding

ancillary information on the vertical extent of clouds

and their liquid water–ice content from collocated sat-

ellite observations, the later is to be called cloudy RO

retrieval. It is emphasized that the issue of ‘‘cloudy RO

retrieval’’ arises only in the data processing step deriving

temperature/humidity profiles from refractivity.

The paper is arranged as follows: section 2 describes

data sources and selection. Section 3 presents a com-

parison of cloud-top temperature among RO ‘‘dry’’ re-

trievals, RO ‘‘wet’’ retrievals, NCEP–NCAR reanalysis,

and ECMWF analysis, as well as model and RO re-

fractivity differences in clear and cloudy air. A new RO

retrieval algorithm is proposed in section 4 for obtaining

in-cloud vertical profiles of temperature from RO re-

fractivity data. The resulting temperature and lapse rate

within clouds using the new retrieval algorithm are com-

pared with those from the standard RO wet retrieval.

Conclusions and future work are provided in section 5.

2. A brief description of observations and analyses

Under the assumption of the spherical symmetry of

the refractive index in the atmosphere, vertical profiles

of bending angle and refractivity can be derived from

the raw RO measurements of the excess Doppler shift of

the radio signals transmitted by GPS satellites (Kursinski

et al. 1997; see also appendix A1 in Zou et al. 1999). The

profiles of refractivity can then be used to generate the so-

called dry and wet retrieval products. For dry air, the

density profiles are first calculated using the relationship

between density and refractivity. The dry temperature

profiles are then derived from the density profiles based

on the hydrostatic equation and the ideal gas law under

the assumption that water vapor partial pressure is zero.

The wet retrieval algorithm estimates both temperature

and water vapor profiles using a 1DVar algorithm (see

Healy and Eyre 2000; Palmer et al. 2000; the algorithm

is described in more detail online at http://cosmic-io.

cosmic.ucar.edu/cdaac/doc/documents/1dvar.pdf).

The COSMIC satellite system consists of a constella-

tion of six low-Earth-orbit (LEO) microsatellites, and

was launched on 15 April 2006 into a circular, 728 in-

clination orbit at 512-km altitude (Anthes et al. 2008).

The first COSMIC GPS RO global datasets of atmo-

spheric parameters (e.g., refractivity, pressure, temper-

ature, etc.) were provided on 21 April 2006. The daily

occultation count was about 400 during the initial 4-month

period and increased after that to about 1400–1600

APRIL 2010 L I N E T A L . 1105

Page 3: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

soundings in August 2006 and up to 2500 more recently.

The vertical resolution ranges from better than 100 m in

the lower troposphere to approximately 0.5 km in the

stratosphere. Each GPS RO measurement quantifies an

integrated refraction effect of the atmosphere on the ray

along the ray path over a few hundred kilometers of space,

centered at the perigee point (Kursinski et al. 1996). The

RO data used in this study are obtained from the UCAR

COSMIC Data Analysis and Archival Center (CDAAC;

Kuo et al. 2004). The CDAAC 1DVar wet retrieval em-

ploys the NCEP–NCAR reanalysis as the first guess field.

The present study uses COSMIC soundings in June 2007

and from June to September 2006.

The cloudy RO soundings were selected based on

CloudSat data. CloudSat was launched into a 705-km

near-circular sun-synchronous polar orbit on 28 April

2006. It orbits Earth approximately once every 1.5 h, fin-

ishing the so-called one observation granule. The primary

FIG. 1. (a) CloudSat orbital track (red curve) at 1702:24 UTC 5 Jun 2007 and a GPS RO sounding (yellow marker) located at (43.798N,

72.988W) whose distance from the CloudSat orbital track is less than 30 km. (b) Reflectivity (unit: dBZ) observed by CloudSat along

a segment of the orbit track indicated by the green line segment in (a). (c) Vertical profile of reflectivity observed by CloudSat at a point on

its track (43.828N, 72.888W), that is closest to the GPS RO location. Dotted lines in (c) indicate the cloud top and cloud base. The yellow

line through the dot in (a) indicates a segment of the tangent line representing the averaged ray direction, the middle point of the line

(yellow dot) indicates the average tangent point position, and the length of the yellow line represents the vertical shift of all the tangent

points of the occultation event.

1106 M O N T H L Y W E A T H E R R E V I E W VOLUME 138

Page 4: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

FIG. 2. (left) Cloud types and (right) vertical profiles of N GPSwet 2 N ECMWF (red solid line) and N GPSwet 2 N NCEP

(blue) for three selected COSMIC cloudy soundings: (a),(b) 0922 UTC 2 Sep 2006; (c),(d) 0331 UTC 20 Aug 2007;

and (e),(f) 2018 UTC 20 Sep 2006. The cloud top and cloud base are indicated by solid black lines. The dashed

horizontal lines indicate the tropopause height calculated from GPS wet retrieval (black), ECMWF (red), and

NCEP–NCAR (blue) analyses. ‘‘Ci’’ stands for cirrus cloud, ‘‘As’’ for altostratus, ‘‘Sc’’ for stratocumulus, and ‘‘Ns’’

stands for nimbostratus cloud.

APRIL 2010 L I N E T A L . 1107

Page 5: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

observing instrument on CloudSat is a 94-GHz, nadir-

pointing cloud profiling radar (CPR), which measures the

returned power backscattered by clouds. The along-track

temporal sample interval equals 0.16 s, resulting in more

than 30 000 vertical profiles of radar reflectivity, liquid

water content, and ice water content for each granule. The

along-track spatial resolution is about 1.1 km, with an

effective field of view (FOV) of approximately 1.4 km 3

2.5 km. Besides reflectivity, liquid water content and ice

water content, cloud layers (with a maximum of five

layers), cloud type, as well as the altitudes of cloud tops

and cloud bases, are also provided by CloudSat (Stephens

et al. 2002).

Cloudy COSMIC RO soundings must collocate with

a CloudSat cloudy profile, where collocation is defined

by a time difference of no more than a half hour and a

spatial separation of less than 30 km. Second, the cloud

top must be above 2 km in order to minimize the impact

of the uncertainty of RO wet retrievals in the lower

troposphere. Third, for simplicity, only soundings with

a single layer cloud are chosen. An example is given in

Fig. 1, which shows the orbital track of an observation

granule initialized at 1702:24 UTC 5 June 2007 and the

location of a nearby RO sounding (Fig. 1a). The radar-

observed reflectivity along a segment on this orbital

track is shown in Fig. 1b and the vertical profile of reflec-

tivity observed by CloudSat at the point on the CloudSat

track that is closest to the RO location is shown in Fig. 1c.

The reflectivity cross section suggests the presence of a

deep cloud system of large horizontal extent.

In the neutral atmosphere the GPS-observed atmo-

spheric refractivity (Nobs) is a function of the pressure (P),

temperature (T), water vapor pressure (Pw), and liquid

water content (W) through the following relationship:

Nobs 5 77.6P

T1 3.73 3 105 P

w

T21 1.4W,

[ Ndry 1 Nwet 1 NLWC, (1)

where P is in millibars, T is in Kelvin, Pw is in millibars,

and W is in liquid water content in grams per cubic me-

ters. The first term on the right-hand side of (1) is referred

to as the dry term, the second one is the wet term, and the

third one is the liquid water content term.

Different types of clouds are usually produced by

different dynamic processes and have different micro-

physical properties (Hartmann et al. 1992; Chen et al.

2000). CloudSat classifies clouds into eight groups (St, Sc,

Cu, Ns, Ac, as, deep convective, and high cloud) based on

structural features of cloud, presence of precipitation,

temperature, and amount of upward radiance (Stephens

et al. 2002). Figure 2 shows vertical profiles of refractivity

differences between GPS RO and large-scale analyses

for three selected COSMIC cloudy RO soundings as

well as cloud types. Large refractivity differences be-

tween NCEP–NCAR reanalysis and ECMWF analysis

and COSMIC observations are seen near the tropo-

pause, cloud top, cloud base, and within clouds.

Vertical profiles of temperature from the ECMWF ana-

lyses along the CloudSat observation granule at CloudSat

data resolution are included in the CloudSat auxiliary data

products and are used for this study. These ECMWF pro-

files are generated from an ECMWF analysis at T799L91

model resolution (see CloudSat Algorithm Process

Description Documents available online at http://www.

cloudsat.cira.colostate.edu/cloudsat_documentation/

html/index.html). [The NCEP–NCAR reanalysis data

used in this study are obtained from the National Oce-

anic and Atmospheric Administration/Earth System Re-

search Laboratory (NOAA/ESRL) online at http://www.

cdc.noaa.gov/data/gridded/data.ncep.reanalysis.spectral.

html at T62L28 model resolution (Kalnay et al. 1996).]

The NCEP–NCAR reanalysis is given on sigma levels,

while COSMIC RO data are given on geometric heights.

To compare the RO refractivity and temperature with

the NCEP–NCAR reanalysis, sigma-level data are first

converted to corresponding geometric heights using the

hydrostatic approximation, the surface pressure, and the

terrain height. Then a horizontal bilinear interpolation is

performed to obtain the temperature at the GPS RO

data location. Finally, a vertical linear interpolation is

FIG. 3. (a) T NCEP (blue), T ECMWF (red), T GPSwet (black), and

T GPSdry (green) at cloud top for the 11 single-layer cloudy sound-

ings in June 2007. (b) Cloud-top height (open circle), cloud types,

and cloud thickness (symbols).

1108 M O N T H L Y W E A T H E R R E V I E W VOLUME 138

Page 6: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

carried out to calculate analysis values at the RO mea-

surement height. Analyses at the 2 times, separated by

a 6-h interval, closest to the RO observation time are

interpolated in time to obtain the final vertical profiles of

the NCEP–NCAR reanalysis that are compared to the

RO data. COSMIC data above 4 km are included in the

June 2007 ECMWF analyses (Healy and Thepaut 2006).

They are not included in the 2006 ECMWF analyses or in

any of the NCEP–NCAR reanalyses.

3. Comparisons between COSMIC GPS RO dataand large-scale analyses

a. Comparison of cloud-top temperatures

To compare temperatures at the cloud top from the

RO standard retrievals with NCEP–NCAR and ECMWF

analyses, we first present (Fig. 3) temperatures at the

cloud top from the dry and wet RO retrievals, NCEP–

NCAR reanalysis, and ECMWF analysis for the 11

single-layer cloudy soundings in June 2007. The cloud-

top height is determined based on CloudSat data, which

are shown in Fig. 3b. The NCEP–NCAR cloud-top tem-

peratures tend to be higher than the other temperatures.

The ECMWF analysis compares more favorably with

the RO retrievals than do the NCEP–NCAR reanalyses.

Moreover, the RO dry temperature retrieval is quite

accurate at the cloud top, except for the cloud whose top

is at 4.3 km. This is expected for the very cold cloud tops

where water vapor makes a negligible contribution to

the RO refractivity. These features characterizing data

for the soundings in June 2007 also hold true for the

cloudy soundings from June to September 2006, which

are included.

Figure 4 plots temperature differences between the

RO, NCEP–NCAR, and ECMWF temperatures at the

cloud-top heights. A systematic warm bias of more

than 1.5 K is found for the NCEP–NCAR temperatures.

The largest warm bias reaches 3.5 K at 9 km, which

could be a result of a combined effect of cold cloud tops

and a cold tropopause at this altitude, both of which may

FIG. 4. Temperature differences at the cloud top: (a) T ECMWF 2 T GPSwet, (b) T NCEP 2

T GPSwet, and (c) T GPSdry 2 T GPSwet. (d) Mean (solid) and RMS differences (dashed) of

T ECMWF 2 T GPSwet (red), T NCEP 2 T GPSwet (blue), and T GPSdry 2 T GPSwet (green). The y axis

indicates cloud-top height.

APRIL 2010 L I N E T A L . 1109

Page 7: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

be inadequately resolved by the model analyses. The

ECMWF temperatures also show a systematic warm

bias at the cloud top in the midtroposphere. The largest

warm bias is about 1.5 K from 3.5 to 5.5 km. In upper

levels, both warm and cold biases are found, but the mag-

nitude is less than 1 K. Because of the omission of water

vapor contribution, a cold bias is expected in the dry re-

trieval compared to the wet retrievals (Fig. 4c). This is also

true for the cloud-top temperature below 5 km in Fig. 4d.

The mean and RMS differences between the ECMWF

and NCEP–NCAR and the RO temperatures near the

cloud top are calculated for three groups of soundings

(Fig. 5): low clouds (cloud-top height is between 2 and

5 km), middle clouds (cloud-top height is between 5 and

8 km), and high clouds (cloud-top height is between 8

and 12 km). Clouds with cloud tops exceeding 12 km are

excluded to avoid the compounded effect of the tropo-

pause. There are 41, 34, and 51 cloudy soundings in low,

middle, and high cloud groups, respectively. The RO

temperatures in the mean tend to be lower than the

model temperatures, with a maximum anomaly of 1–2 K

near and just below the cloud top. The consistency in the

shape of the profiles suggests that the RO soundings are

measuring real cloud effects that are not in the analyses,

particularly the NCEP–NCAR reanalysis.

b. Model and RO refrectivity differences in clearand cloudy air

Earlier studies have found a negative N bias in the

lower and midtroposphere (Rocken et al. 1997; Ao et al.

2003; Sokolovskiy 2003; Hajj et al. 2004; Beyerle et al.

2006; Xie et al. 2006; Anthes et al. 2008). Comparing

CHAMP observations with ECMWF analysis, Hajj et al.

(2004) documented that the negative fractional refrac-

tivity difference is nearly 1%–4% below 10 km. Anthes

et al. (2008) found that negative N bias only occurs be-

low 3 km for COSMIC with a magnitude of ;1%. These

studies did not distinguish between clear and cloudy air.

Here we consider whether the mean differences be-

tween models and RO vary between clear and cloudy

air. The total number of cloudy soundings is increased to

131 by including multilayer clouds. A total of 86 clear-sky

FIG. 5. Mean (solid) and RMS (dashed) of T ECMWF 2 T GPSwet (red) and T NCEP 2 T GPSwet

(blue) within 2 km above and below the cloud top for soundings with the cloud top between (a)

2 and 5, (b) 5 and 8, and (c) 8 and 12 km. (d) The total numbers of cloudy soundings included in

(a) (solid), (b) (dashed), and (c) (dotted).

1110 M O N T H L Y W E A T H E R R E V I E W VOLUME 138

Page 8: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

RO soundings are identified with collocated CloudSat

data based on the same criteria for the identification of

cloudy profiles that are described in section 2. Figure 6

presents a comparison between RO and the ECMWF

and NCEP–NCAR refractivities in cloudy- and clear-sky

conditions. For cloudy soundings, a positive N difference

(NGPS . Nanalysis) is found throughout the entire layer of

the atmosphere below 18 km. This positive difference

may be caused by high water vapor content as well as the

presence of liquid water in the clouds, which would tend

to increase the observed RO N [see Eq. (1) in section 2].

In contrast, a negative N difference (NGPS , Nanalysis) is

found for clear-sky soundings. The magnitude of negative

clear-sky bias is slightly smaller than the positive cloudy

bias. The RMS differences are also smaller under clear-

sky conditions.

4. A new GPS cloudy retrieval algorithm

a. Cloudy retrieval assuming saturation

The new cloudy retrieval starts from the cloud top

(Ztop) and goes downward to the cloud base (Zbase), where

Ztop and Zbase are obtained from CloudSat. Cloud-top

temperature (T0) and pressure (P0) are defined as the

RO wet temperature and pressure at cloud top (Ztop).

The values of Ztop, T0, and P0 serve as the upper bound-

ary conditions for the cloudy profile retrieval.

For saturated cloudy air, the equation of state can be

written as

P 5 rRdT(1 1 0.61 3 q

s), (2)

where Rd is the gas constant for dry air (Rd 5 287

J kg21 K21), T is the temperature, and qs is the satura-

tion specific humidity.

The saturation specific humidity qs in (2) can be ex-

pressed as a function of the saturation vapor pressure (es):

qs5

0.622 3 Pw

P� 0.378 3 Pw

50.622 3 e

s(T)

P� 0.378 3 es(T)

, (3)

which is a function of temperature (T) only and

es(T) 5 6.112 exp

17.67(T � 273.15)

T � 29.65

� �. (4)

FIG. 6. (left) Mean and (right) RMS around the mean of COSMIC GPS and ECMWF

fractional refractivity difference (a),(b) (N GPS 2 N ECMWF)/N ECMWF and (c),(d) (N GPS 2

N NCEP)/N NCEP for cloudy soundings (solid) and clear soundings (dashed).

APRIL 2010 L I N E T A L . 1111

Page 9: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

In (4), es is in hectopascals and T is in Kelvin. It is an

expression with accuracy of 0.3% for jTj# 358C (Bolton

1980).

Substituting (3) into (2), we obtain

r 5P

RdT 3 1 1 0.61 3

0.622 3 es(T)

P� 0.378 3 es(T)

� � [ r(P, T).

(5)

Therefore, density (r) within a saturated cloud is a func-

tion of temperature (T) and pressure (P) alone.

Starting from the cloud top, the following procedure

can be followed to obtain the in-cloud profiles of pres-

sure and temperature from the refractivity observations.

We use the subscript ‘‘m’’ to denote the vertical level,

and m 5 0 to represent the cloud top. The hydrostatic

equation can be integrated to obtain the pressure at the

(m 1 1)th vertical level below the mth level:

Pm11

5 Pm� g

mr

m(z

m11� z

m), where

rm

5 rm(P

m, T

m), (6)

and where gm is the gravity constant with its altitude

dependence accounted for (Glickman 2000). It is noticed

that a constant density rm is used within the half-open

interval zm 1 1 , z # zm in (6), which may introduce a

small deviation from the hydrostatic balance. The tem-

perature at the (m 1 1)th vertical level (TGPSsatm11 ) can be

obtained from (1) given refractivity observations (Nobsm11)

at the (m 1 1)th vertical level and pressure (Pm11) cal-

culated by (6). The effect of liquid water term on the RO

temperature retrieval is negligible as will be shown later

using CloudSat observations for Wm11 within clouds.

The value of TGPSsatm11 is determined to be within the

temperature range [Tm 2 58C, Tm 1 58C]. The squared

differences of refractivity between the GPS RO ob-

servations (Nobsm11) and model refractivity (N) are cal-

culated at different temperatures in this temperature

range at 0.18C interval. The temperature at which

J(T) 5 [N(T)�Nobsm11]2 reaches minimum is taken as

the value for TGPSsatm11 [i.e., J(TGPSsat

m11 ) 5 min J(Tm11)].

Figure 7 illustrates this process for obtaining TGPSsatm11 for

the RO sounding located at 67.228S, 144.18W on 1437

UTC 21 July 2006. The distributions of J(Ti) (i 5 1,

2, . . . , 100) at six selected vertical levels are shown by

black curves in Fig. 7. The solutions of cloudy retrieval

TGPSsatm11 assuming saturation are the minimum points on

these curves [i.e., J(TGPSsatm11 ) 5 min

TJ(T)]. The purple

line connecting all the minimum points (at different

vertical levels) gives the vertical profile of RO temper-

ature assuming saturation (TGPSsat).

To show how sensitive GPS cloudy retrieval is to the

saturation assumption, we show in Fig. 8 the differ-

ences between the RO temperature retrievals assuming

FIG. 7. The squared difference of refractivity (black curve) be-

tween GPS RO observations (N obs) and model-calculated re-

fractivity at different temperatures, assuming saturation within

cloud (N sat) at six selected vertical levels (altitudes are indicated on

the right). The solution of GPS saturation retrieval T GPSsat where

N obs 2 N sat ’ 0 is represented by the purple line. The RO sound-

ing shown in this figure is located at 67.228S, 144.18W at 1437 UTC

21 Jul 2006.

FIG. 8. Dependence of the differences between GPS saturation

retrieval and wet retrieval (TGPSsat 2 TGPSwet) on relative humidity

calculated from ECMWF analysis using data from the middle of

the cloud.

1112 M O N T H L Y W E A T H E R R E V I E W VOLUME 138

Page 10: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

saturation (TGPSsat) and the standard GPS wet retrieval

(TGPSwet) at the middle height of the cloud. As expected,

the difference between TGPSsat and TGPSwet is quite small

when the relative humidity is nearly 100% (Fig. 8). The

temperature difference is ,648C when the relative hu-

midity is greater than 85%. However, differences of

.648C are found for soundings with relative humidity

less than 85% (Fig. 8).

b. An empirical GPS RO cloudy retrieval algorithmwithout assuming saturation

Saturated clouds usually occupy only a fraction of the

area over which the averaged state of the atmosphere is

either measured by the RO technique or represented by

the ECMWF or NCEP–NCAR analyses. Therefore,

similar to what is often done in NWP models in which

convection is initiated when the relative humidity ex-

ceeds a threshold value (such as 85% instead of 100%),

we introduce an empirical parameter, a, in the retrieval

of the RO temperature within cloud. The observed GPS

refractivity is expressed as a weighted function of clear

and cloudy refractivities:

Nobssm11 5 (1� a)Nclear

m11 1 aNcloudm11 , (7)

where Nclearm11 5 N

drym11 1 Nwet

m11, Ncloudm11 5 N

drym11 1 Nsat

m11,

and Nsat 5 3.73 3 105(es/T2). The temperature at the

(m 1 1)th level is obtained as a solution to the above

equation. The saturation assumption corresponds to a

totally cloudy condition (i.e., a 5 1).

FIG. 9. Vertical variations of (a) liquid water content and (b) ice water content observed by CloudSat within the

cloud. Vertical mean of (c) liquid water content and (d) ice water content. The y axis in (c),(d) is the vertical mean

of relative humidity within cloud calculated from ECMWF analysis. The dashed line in (d) represents a linear

regression model between ice water content and relative humidity with cloud. The open circles in (d) indicates

outlier removed from linear fitting whose distance to the biweighting mean is more than 3 times the biweighting

standard deviation.

APRIL 2010 L I N E T A L . 1113

Page 11: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

The mean in-cloud relative humidity calculated from

ECMWF analysis for the 147 clouds is in the range of

60%–80%. The scatter (RMS difference from the mean)

for the NCEP–NCAR reanalysis is more than twice as

large as for ECMWF or GPS wet retrieval. The de-

pendence of the relative humidity on the liquid–ice

water content is shown in Fig. 9. Because of missing data

from CloudSat, a total of 19 liquid water clouds and 67

ice water clouds are found. Surprisingly, we find no re-

lationship between the observed liquid water content

and the ECMWF relative humidity (RH). For ice water

clouds, however, a linear relationship between ECMWF

RH and ice water content is found (Fig. 9d). Having

removed outliers (a total of 13) whose distances to the

mean are greater than 3 times of the standard deviation

(Lanzante 1996) and by assuming a 5 RH, a linear re-

gression model is developed for determining the pa-

rameter a from the observed IWC:

a 55.273 3 IWC 1 0.6849, IWC # 0.05975 g m�3

1 IWC . 0.05975 g m�3,

(

(8)

where IWC is the vertically averaged ice water content.

Then, the formulation (8) for a is applied to those 67 ice

clouds for determining the in-cloud temperature. For

liquid water clouds, the average value of relative hu-

midity is used for a (50.8).

Considering the uncertainty of the initial con-

dition (T0 and P0) at the cloud top, we set the

initial condition for in-cloud profile retrieval as

T0

5 TGPSwet � (TGPSwet � TECMWF) 6 (i/2)sT

, P0

5

PGPSwet � (PGPSwet � PECMWF) 6 (i/2)sP

(i 5 1, 2),

where sT2 and sP

2 are variances of the temperature and

pressure of the GPS wet retrieval at the cloud top and are

estimated from the differences between the GPS wet re-

trieval and the ECMWF analysis. The final retrieval is

defined as the mean of the temperature profiles obtained

using nine perturbed boundary conditions at the cloud top.

c. Numerical results

The contributions of liquid water content to the total

refractivity are first examined for the 19 water clouds for

which CloudSat data is available and the maximum

liquid water content ranges from 0.05 to 1.4 g m23.

Figure 10 presents the fractional refractivity differences

between COSMIC GPS and ECMWF (NGPS 2 NECMWF)/

NECMWF with and without including cloud liquid water

term. First, COSMIC GPS and ECMWF fractional re-

fractivity differences are positive for all water clouds.

Second, within all liquid clouds, fractional refractivity

differences between COSMIC GPS and ECMWF with

liquid water contribution included are consistently larger

than those without. In other words, if the presence of

hydrometeors is disregarded in the analysis of RO data, it

will result a positive bias. Third, contribution of liquid

water to GPS observed refractivity is not negligible unless

the liquid water content is less than 0.05 g m23. These

results deduced from water clouds are consistent with the

sign of the N bias seen in Fig. 6 for cloudy ROs.

Differences of temperature retrievals within clouds

using our method and the standard RO wet retrieval

algorithm (TGPScloud 2 TGPSwet) are shown in Fig. 11.

Mean differences are calculated using the biweight es-

timation method (Lanzante 1996; Zou and Zeng 2006).

Data points whose distance to the mean is more than 3

times of the standard deviation at any vertical level are

identified as outliers. This leads to a total of 74 cloudy

soundings that have liquid water/ice water data from

CloudSat. The liquid water contribution to refractivity

[i.e., the third term in (1)] is included in the retrieval.

The mean differences are calculated by aligning all

clouds by cloud middle height (Figs. 11a,b), cloud top

(Figs. 11c,d) and cloud base (Figs. 11e,f). It is found that

the cloudy retrievals produced a warmer temperature

within clouds, but a colder temperature near the cloud

base than the RO wet retrievals. The differences be-

tween the cloudy and wet retrievals are relatively small

near the cloud top. The largest positive bias is about 2 K,

which occurred at about 1.5 km above the middle of the

cloud. The largest negative bias is at the cloud base, with

a magnitude of 1.5 K. The 99% confidence interval is

less than 0.5 K.

FIG. 10. COSMIC GPS and ECMWF fractional refractivity dif-

ference (N GPS 2 N ECMWF)/N ECMWF with (stars) and without

(dots) including the cloud liquid water term at the height of the

maximum liquid water content (color bar) within 19 water clouds

for which CloudSat data are available.

1114 M O N T H L Y W E A T H E R R E V I E W VOLUME 138

Page 12: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

The effect of cloudy retrieval on the lapse rate within

clouds is shown in Fig. 12. It is seen that the tempera-

ture lapse rates are close to moist lapse rates that vary

from 5.6 to 6.8 K km21 on average (see Table 1). Also

shown in Fig. 12 is a simple cloudy retrieval using

a constant value 0.85 for the parameter a (to be denoted

T GPScloud85) instead of the linear regression model in (8)

for all ice clouds. At the middle of cloud, the two cloudy

retrievals agree well with ECMWF profiles, giving a

lapse rate of about 6.08C km21. The GPS wet retrieval

produces a slightly larger lapse rate (5.98C km21) than

the cloudy retrievals and that of the NCEP–NCAR re-

analysis is as small as 5.48C km21. The lapse rate from

the RO cloudy retrievals increases from cloud middle to

cloud top, reaching a value of about 7.58C km21. The

lapse rate from the ECMWF analysis and NCEP–NCAR

reanalysis first increases with height from the cloud mid-

dle, then decreases with height when approaching the

FIG. 11. Vertical profiles of the mean and the 99% confidence interval (dashed line) of

temperature differences between the cloudy retrievals and the RO wet retrieval [i.e., T GPScloud 2

T GPSwet (green) and T GPScloud85 2 T GPSwet (purple)] within clouds aligned at (a),(b) the middle

of the cloud; (c),(d) the cloud top; and (e),(f) the cloud base.

APRIL 2010 L I N E T A L . 1115

Page 13: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

cloud top. The mean lapse rate derived from the GPS

standard wet retrieval is smaller than that from the

ECMWF analysis but larger than that from NCEP–

NCAR reanalysis. In the lower part of the cloud, the

mean lapse rate from all data sources does not change

greatly with height, again with ECMWF analysis com-

paring more favorably with the cloudy retrievals than

the NCEP–NCAR reanalysis and GPS standard wet

retrieval. The NCEP–NCAR lapse rate is larger than the

ECMWF lapse rate and the GPS wet retrieval is even

larger than the NCEP–NCAR lapse rate. The mean

lapse rate derived from a cloudy retrieval with a 5 0.85

for ice clouds is qualitatively similar to results from the

cloudy retrieval that requires ice water content as input.

FIG. 12. (left) Biweight mean and (right) standard deviation of the lapse rate within the

cloud calculated from TGPSwet (black), TECMWF (red), TNCEP (blue), TGPScloud (green), and

TGPScloud85 (purple) aligned by (a),(b) the middle height of cloud; (c),(d) the cloud top; and

(e),(f) the cloud base. The numbers on the y axis indicate the vertical distance above (positive)

and below (negative) the middle height, the cloud-top height, and the cloud-base height.

1116 M O N T H L Y W E A T H E R R E V I E W VOLUME 138

Page 14: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

5. Summary

In this study we examined the vertical structure of

temperatures and lapse rates within clouds as retrieved

from COSMIC GPS RO refractivity profiles and com-

pared to temperature profiles from ECMWF and NCEP–

NCAR analyses. CloudSat data were used to determine

the presence or absence of clouds at the location of the

RO observations. GPS RO provides vertical profiles of

temperature and water vapor in cloudy regions over the

globe with high vertical resolution.

By separating cloudy- and clear-sky COSMIC RO

soundings, we find that the observed RO refractivities

N are greater on average than the refractivities of the

ECMWF and NCEP–NCAR large-scale analyses within

clouds [i.e., (NGPS �Nanalysis)/Nanalysis . 0, where the

bar represents the average over many cloudy ROs,

throughout the atmosphere (from 2 to 18 km)]. In con-

trast, the observed RO refractivities in clear areas are

less in the mean than the refractivities in the ECMWF

and NCEP–NCAR analyses.

The standard GPS RO wet temperature retrieval used

in previous studies does not distinguish between cloudy-

and clear-sky conditions. A new cloudy retrieval algorithm

is proposed in this study. By introducing a sounding-

dependent relative humidity parameter a that depends

on the ice water content, the in-cloud temperature

profile is retrieved from a weighted sum of clear-sky and

cloudy refractivity values. Compared to the standard wet

retrieval, the cloudy temperature retrieval is consistently

warmer within clouds by ;2 K and slightly colder near

the cloud top and cloud base (1–2 K), leading to a more

rapid increase of the lapse rate with height in the upper

half of the cloud than in both the ECMWF and NCEP–

NCAR analyses. The average lapse rate within clouds is

nearly constant and close to the moist lapse rate in the

lower half of the clouds. In the absence of ice water

content measurements, an empirical value of 85% seems

to be a good approximation to a.

CloudSat data is used in this study to develop a linear

regression model for a moisture parameter incorporated

in the cloudy RO retrieval and to examine the sensitivity

of the cloudy retrieval to hydrometeor concentration

within clouds. The linear regression model that was

developed in this study could (i) be applied to the cloudy

RO retrieval using model-predicted hydrometeor con-

centration if collocated observations of hydrometeor

concentrations are not available and (ii) be used in the

observation operator for assimilation of refractivity within

clouds to account for the effect of cloudiness.

Future work includes extending the present study to

include a longer period of data, validation, and impact

study. It is anticipated that these in-cloud profile data

will be extremely valuable for providing a better un-

derstanding of the in-cloud atmospheric thermodynamics

structures, physical parameterization of moist processes,

and global radiative energy budget. The in-cloud vertical

variations of the atmospheric thermodynamic state will

also be extremely valuable for the assimilation of cloudy

radiance data, which are yet to be used by most opera-

tional centers.

Acknowledgments. This research is supported by

NOAA/FIU Grant 12000586-09.

REFERENCES

Anthes, R. A., C. Rocken, and Y. H. Kuo, 2000: Applications of

COSMIC to meteorology and climate. TAO, 11, 115–156.

——, and Coauthors, 2008: The COSMIC/FORMOSAT-3 mission:

Early results. Bull. Amer. Meteor. Soc., 89, 313–333.

Ao, C. O., T. K. Meehan, G. A. Hajj, A. J. Mannucci, and G. Beyerle,

2003: Lower troposphere refractivity bias in GPS occultation

retrievals. J. Geophys. Res., 108, 4577, doi:10.1029/2002JD003216.

Bayler, G. M., R. M. Aune, and W. H. Raymond, 2000: NWP cloud

initialization using GOES sounder data and improved model-

ing of nonprecipitating clouds. Mon. Wea. Rev., 128, 3911–

3920.

Beyerle, G., T. Schmidt, J. Wickert, S. Heise, M. Rothacher, G. Konig-

Langlo, and K. B. Lauritsen, 2006: Observations and simulations

of receiver-induced refractivity biases in GPS radio occultation.

J. Geophys. Res., 111, D12101, doi:10.1029/2005JD006673.

Bolton, D., 1980: The computation of equivalent potential tem-

perature. Mon. Wea. Rev., 108, 1046–1053.

Chen, T., W. B. Rossow, and Y. Zhang, 2000: Cloud type radiative

effects from the international satellite cloud climatology pro-

ject. Preprints, 11th Symp. on Global Change Studies, Long

Beach, CA, Amer. Meteor. Soc., 86–89.

Diak, G. R., M. C. Anderson, W. L. Bland, J. M. Norman, J. M.

Mecikalski, and R. M. Aune, 1998: Agricultural management

decision aids driven by real-time satellite data. Bull. Amer.

Meteor. Soc., 79, 1345–1355.

Frey, R. A., B. A. Baum, W. P. Menzel, S. A. Ackerman, C. C.

Moeller, and J. D. Spinhirne, 1999: A comparison of cloud top

heights computed from airborne lidar and MAS radiance data

using CO2 slicing. J. Geophys. Res., 104, 24 547–24 555.

Glickman, T. Ed., 2000: Glossary of Meteorology. 2nd ed. Amer.

Meteor. Soc., 855 pp.

Gorbunov, M. E., and L. Kornblueh, 2001: Analysis and validation

of GPS/MET radio occultation data. J. Geophys. Res., 106

(D15), 17 161–17 169.

TABLE 1. Biweight mean and standard deviation (in parenthe-

ses) of the moist-adiabatic lapse rate calculated from GPS wet

retrieval, ECMWF analysis, NCEP–NCAR analysis, and GPS RO

cloud retrieval data at the middle height of the cloud, cloud top,

and cloud base. Unit: K km21.

GPS wet ECMWF NCEP Cloud retrieval

Cloud middle 5.8 (1.0) 5.7 (1.1) 5.8 (1.0) 5.9 (1.2)

Cloud top 6.2 (0.6) 6.4 (0.7) 5.9 (0.8) 6.8 (0.5)

Cloud base 5.8 (1.4) 5.8 (1.3) 5.6 (1.4) 5.6 (1.6)

APRIL 2010 L I N E T A L . 1117

Page 15: COSMIC GPS Radio Occultation Temperature Profiles in Clouds

Hajj, G. A., and Coauthors, 2004: CHAMP and SAC-C atmospheric

occultation results and intercomparisons. J. Geophys. Res., 109,

D06109, doi:10.1029/2003JD003909.

Hartmann, D. L., M. E. Ockert-bell, and M. L. Michelsen, 1992:

The effect of cloud type on Earth’s energy balance: Global

analysis. J. Climate, 5, 1281–1304.

Hawkinson, J. A., W. Feltz, and S. A. Ackerman, 2005: A com-

parison of GOES sounder- and cloud lidar- and radar-retrieved

cloud-top heights. J. Appl. Meteor., 44, 1234–1242.

Healy, S., and J. Eyre, 2000: Retrieving temperature, water vapor

and surface pressure information from refractivity-index pro-

files derived by radio occultation: A simulation study. Quart. J.

Roy. Meteor. Soc., 126, 1661–1683.

——, and J.-N. Thepaut, 2006: Assimilation experiments with

CHAMP GPS radio occultation measurements. Quart. J. Roy.

Meteor. Soc., 132, 605–623.

Hollars, S., Q. Fu, J. Comstock, and T. Ackerman, 2004: Com-

parison of cloud-top height retrievals from ground-based 35

GHz MMCR and GMS-5 satellite observations at ARM TWP

Manus site. Atmos. Res., 72, 169–186.

Holz, R. E., S. Ackerman, P. Antonelli, F. Nagle, R. O. Knuteson,

M. McGill, D. L. Hlavka, and W. D. Hart, 2006: An im-

provement to the high-spectral-resolution CO2-slicing cloud-

top altitude retrieval. J. Atmos. Oceanic Technol., 23, 653–670.

Kahn, B. H., A. Eldering, A. J. Braverman, E. J. Fetzer, J. H. Jiang,

E. Fishbein, and D. L. Wu, 2007: Toward the characterization

of upper tropospheric clouds using Atmospheric Infrared

Sounder and Microwave Limb Sounder observations. J. Geo-

phys. Res., 112, D05202, doi:10.1029/2006JD007336.

Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year Re-

analysis Project. Bull. Amer. Meteor. Soc., 77, 437–471.

Kishore, P., S. P. Namboothiri, J. H. Jiang, V. Sivakumar, and

K. Igarashi, 2008: Global temperature estimates in the tro-

posphere and stratosphere: A validation study of COSMIC/

FORMOSAT-3 measurements. Atmos. Chem. Phys. Discuss.,

8, 8327–8355.

Kursinski, E. R., and Coauthors, 1996: Initial results of radio oc-

cultation observations of Earth’s atmosphere using the global

positioning system. Science, 271, 1107–1110.

——, G. A. Hajj, J. T. Schofield, R. P. Linfield, and K. R. Hardy,

1997: Observing Earth’s atmosphere with radio occultation

measurements using the Global Positioning System. J. Geo-

phys. Res., 102, 23 429–23 465.

Kuo, Y.-H., T.-K. Wee, S. Sokolovskiy, C. Rocken, W. Schreiner,

D. Hunt, and R. A. Anthes, 2004: Inversion and error esti-

mation of GPS radio occultation data. J. Meteor. Soc. Japan,

82, 507–531.

Lanzante, J., 1996: Resistant, robust and nonparametric techniques

for the analysis of climate data: Theory and examples, in-

cluding applications to historical radiosonde station data. Int.

J. Climatol., 16, 1197–1226.

Mahesh, A., M. A. Gray, S. P. Palm, W. D. Hart, and J. D. Spinhirne,

2004: Passive and active detection of clouds: Comparisons be-

tween MODIS and GLAS observations. Geophys. Res. Lett.,

31, L04108, doi:10.1029/2003GL018859.

Naud, C., J.-P. Muller, and E. E. Clothiaux, 2002: Comparison of

cloud top heights derived from MISR stereo and MODIS CO2-

slicing. Geophys. Res. Lett., 29, 1795, doi:10.1029/2002GL015460.

——, ——, M. Haeffelin, Y. Morille, and A. Delaval, 2004: As-

sessment of MISR and MODIS cloud top heights through inter-

comparisons with a back-scattering lidar at SIRTA. Geophys.

Res. Lett., 31, L04114, doi:10.1029/2003GL018976.

Palmer, P. I., J. Barnett, J. Eyre, and S. Healy, 2000: A non-linear

optimal estimation inverse method for radio occultation

measurements of temperature, humidity, and surface pres-

sure. J. Geophys. Res., 105, 17 513–17 526.

Rocken, C., and Coauthors, 1997: Analysis and validation of GPS/

MET data in the neutral atmosphere. J. Geophys. Res., 102,

29 849–29 866.

Sherwood, S. C., J.-H. Chae, P. Minnis, and M. McGill, 2004: Un-

derestimation of deep convective cloud tops by thermal imagery.

Geophys. Res. Lett., 31, L11102, doi:10.1029/2004GL019699.

Smith, W. L., and C. M. R. Platt, 1978: Comparison of satellite-

deduced cloud heights with indications from radiosonde and

ground-based laser measurements. J. Appl. Meteor., 17, 1796–

1802.

——, and R. Frey, 1990: On cloud altitude determinations from

high resolution interferometer sounder (HIS) observations.

J. Appl. Meteor., 29, 658–662.

Sokolovskiy, S., 2003: Effect of superrefraction on inversions of

radio occultation signals in the lower troposphere. Radio Sci.,

38, 1058, doi:10.1029/2002RS002728.

Stephens, G. L., S. C. Tsay, J. P. W. Stackhouse, and P. Flatau,

1990: The relevance of the microphysical and radiative prop-

erties of cirrus clouds to climate and climatic feedback.

J. Atmos. Sci., 47, 1742–1753.

——, and Coauthors, 2002: The CloudSat mission and the A-train:

A new dimension of space-based observations of clouds and

precipitation. Bull. Amer. Meteor. Soc., 83, 1771–1790.

Stubenrauch, C. J., F. Eddounia, and L. Sauvage, 2005: Cloud

heights from TOVS Path-B: Evaluation using LITE observa-

tions and distributions of highest cloud layers. J. Geophys.

Res., 110, D19203, doi:10.1029/2004JD005447.

Ware, R., and Coauthors, 1996: GPS sounding of the atmosphere

from low earth orbit: Preliminary results. Bull. Amer. Meteor.

Soc., 77, 19–40.

Weisz, E., J. Li, J. Li, D. K. Zhou, H.-L. Huang, M. D. Goldberg, and

P. Yang, 2007a: Cloudy sounding and cloud-top height retrieval

from AIRS alone single field-of-view radiance measurements.

Geophys. Res. Lett., 34, L12802, doi:10.1029/2007GL030219.

——, ——, W. P. Menzel, A. K. Heidinger, B. H. Kahn, and C.-Y. Liu,

2007b: Comparison of AIRS, MODIS, CloudSat and CALIPSO

cloud top height retrievals. Geophys. Res. Lett., 34, L17811,

doi:10.1029/2007GL030676.

Wickert, J., and Coauthors, 2001: Atmosphere sounding by GPS

radio occultation: First results from CHAMP. Geophys. Res.

Lett., 28, 3263–3266.

——, T. Schmidt, G. Beyerle, R. Konig, C. Reigber, and

N. Jakowski, 2004: The radio occultation experiment aboard

CHAMP: Operational data analysis and validation of vertical

atmospheric profiles. J. Meteor. Soc. Japan, 82, 381–395.

Wylie, D. P., and P.-H. Wang, 1999: Comparison of SAGE-II and

HIRS co-located cloud height measurements. Geophys. Res.

Lett., 26, 3373–3375.

Xie, F., S. Syndergaard, E. R. Kursinski, and B. M. Herman, 2006:

An approach for retrieving marine boundary layer refractivity

from GPS occultation data in the presence of superrefraction.

J. Atmos. Oceanic Technol., 23, 1629–1644.

Zou, X., and Z. Zeng, 2006: A quality control procedure for GPS radio

occultation data. J. Geophys. Res., 111, D02112, doi:10.1029/

2005JD005846.

——, and Coauthors, 1999: A ray-tracing operator and its adjoint

for the use of GPS/MET refraction angle measurements.

J. Geophys. Res., 104 (D18), 22 301–22 318.

1118 M O N T H L Y W E A T H E R R E V I E W VOLUME 138