Transcript Document

TURBULENCE AND MICROPHYSICAL RETRIEVALS
USING MM-WAVELENGTH DOPPLER SPECTRA
Pavlos Kollias, Bruce A. Albrecht and Benjamin J. Dow
Rosenstiel School of Marine and Atmospheric Science (RSMAS),
University of Miami
Millimeter wavelength Doppler radars operating at 94
GHz (e.g. Lhermitte 1987); have high temporal and
spatial resolution, extreme sensitivity, and high velocity
resolution. Due to their short wavelength, millimeter
radars are capable of detecting very small droplets. This
capability of millimeter wave radars has established them
as the ultimate observing systems for the study of weak
meteorological targets undetectable by conventional
weather radars. Parallel with the development of
millimeter wavelength radars, retrieval techniques for the
estimation of cloud and drizzle drop distributions were
developed (Gossard 1994; Frisch et al., 1995; Babb et
al., 1999). In this paper, some new ideas on the use of
millimeter wavelength Doppler spectra for the retrieval of
turbulence and microphysical parameters are considered
in an attempt to emphasize the importance of recording
the Doppler spectrum.
In the case of fair-weather cumuli and non-drizzling stratus layers, cloud droplets are the main source of backscattering. The terminal velocity of a cloud droplet is small, (0.3
cm s-1 and 7 cm s-1 for a 10µm and a 50µm droplet, respectively) so that the droplets’ vertical velocity is primarily due to air motion and turbulence. The cloud droplets’ inertia
is small so they are good tracers of turbulent air velocity in the same way that smoke particles reveal turbulent eddies in a smoke filled room. The time-height mapping of the
mean Doppler velocity reveals the cloud internal circulation structure in terms of updrafts-downdrafts (Kollias and Albrecht, 2000; Kollias et al., 2001) that can use to analyze
the internal dynamics and the interaction between the cloud and its environment. Examples of Large-Eddy Observations (LEO, Figs. 1-3), power spectra of vertical velocity
time series (Fig. 4) and calculations of fractional area of updrafts and downdrafts (Fig. 5) using direct sampling or the statistics of the vertical velocity (Randall et al., 1992) in
a non-drizzling marine stratus are shown in the above figures.
A variation of 2 ms-1 across the beam
cross section, which is likely to happen in
a fair weather cumulus, will create a very
large spectrum variance. Fair-weather
cumuli are highly turbulent. We often
observe Doppler spectrum bimodality that
indicates the presence of sharp vertical
velocity gradients such as those in the
region between adjacent updrafts and
downdrafts (Albrecht et al, 2001).
For a typical cloud droplet distribution, the
expected spread of the droplets’ terminal
velocities (0.3 to 10 cms-1) and the
associated Doppler spectrum variance will be
only a few cm2s-2. As a result the observed
Doppler spectrum width arises from
turbulence and systematic variation of the
vertical wind across the beam.
Depending on the character of the systematic
variation of w across the beam, either wide or
bimodal spectra can be produced.
Fig. 6: Mean Doppler velocity estimates using FFT
calculated using raw I/Q time series in a cirrus cloud.
2,500 I/Q pairs were used for each FFT. The data
were collected within (black) and above (red) a cirrus
fallstreak. Notice the high-resolution mean Doppler
variability within the fallstreak caused by microshear.
Fig. 7: Examples of FFTs (top) calculated using
raw I/Q time series (2,500 pairs – 0.5 sec dwell
time). The data correspond to the time period 2630 sec in Fig. 6. Bottom: The resulting Doppler
spectra if all the I/Q data were used for the
calculation of a single FFT. Notice the spectra
broadening due to the linear variation of w.
Frisch et al., (1995) demonstrate a retrieval technique for drizzle
microphysical parameters using the Doppler moments. Assuming a
lognormal drizzle size distribution, the parameters (N-concentration, romodal radius and x-lognormal distribution width) are retrieved by the first
three Doppler moments. While the assumption that the drizzle distribution
controls reflectivity and mean Doppler velocity is plausible, especially if
conditional sampling (threshold values for dBZ and mean Doppler) is
applied to the data set, this is not clear for the Doppler spectrum width.
The width of the lognormal size distribution is parameterized as the ratio
of the second and first moment of the Doppler spectrum. Data from the
Drizzle and Entrainment Cloud Study 99 (DECS’99) experiment show
that the turbulence contribution to the second moment is large and has to
be removed. Otherwise, x and N will be significantly overestimated and ro
underestimated. Fig. 1 shows time series of standard deviation of mean
Doppler velocity (dashed line) and Doppler spectrum width (solid line).
The variance of the mean Doppler velocity 2w is an indicator of
turbulence intensity at scales larger than the sampled volume, while the
Doppler spectrum variance 2 is an indicator of turbulence intensity within
the sampled volume. During non-drizzling periods the main contributor to
Doppler spectrum width is turbulence and thus t = Rw (R1 in our
case). In general, the ratio R = t/w can be estimated theoretically from
the Kolmogorov turbulence theory. During drizzling periods:
Fig. 8: Simulation of Doppler spectra for different
types (linear (top left) and step-function (bottom
left)) of sub-resolution volume cross-radial wind
variation. Linear wind fields result in Doppler
spectrum
broadening.
Discontinuous
(step
function) wind field profile results in bimodal
Doppler spectrum.
Fig. 9: Time-height mapping of Doppler spectrum
width in a fair-weather cumulus. The white contour
represents the 1.5 ms-1 contour identified as the
updraft boundary. Note the large spectrum width
values near the updraft boundary.
Therefore we need to correct the Doppler spectrum width for turbulence broadening.
This is done by using the w during the drizzling period and the estimate of R from
non-drizzling periods. In other words, we assume that the variance due to the drizzle
drop size distribution (DSD) can be estimated from the difference between the
Doppler spectrum variance and the mean Doppler velocity variance at the same
altitude:
2
2
 DSD    ( R w )
The lognormal size distribution width x is given at the ratio of the DSD and the mean
Doppler velocity (Frisch et al., 1995). Our modification to the original Frisch technique
can lead to more realistic microphysical retrievals since the turbulence broadening is
essentially removed. Finally, higher Doppler spectra moments can be incorporated
into the retrieval algorithm. The data have showed a remarkable correlation between
the mean Doppler velocity and the Doppler spectrum skewness. Although turbulence
acts as a smearing mechanism and partially removes the asymmetry of the Doppler
spectrum caused by the large drops velocity tail, the skewness of the Doppler
spectrum can be expressed as a function of the lognormal size distribution
parameters and therefore can be used to enhance the robustness of the retrievals.
Fig. 10 Top: Cross section of mean Doppler
velocity (blue) and Doppler spectrum width at
1.55 km altitude within the fair weather cumuli.
Note the large values of spectrum width
associated with large wind shear zones at the
updraft boundaries. Bottom: Example of bimodal
spectra near the updraft boundaries.
The recorded Doppler spectra from millimeter wavelength
radars can be used to obtain critical information on
microphysical processes and their interaction with updraft
and downdraft structures. As the retrieval techniques for
cloud radar become more sophisticated, it will be possible
to further advance our understanding by providing critical
observations for process studies and the direct evaluation
of cloud models.
Albrecht, B. A., P. Kollias and B. J. Dow, 2001. Millimeterwavelength radar observations of updrafts, downdrafts and
turbulence in fair weather cumuli, 30th International
Conference on Radar Meteorology, Munich, Germany, 1924 July 2001.
Babb, D., J. Verlinde and B.A. Albrecht, 1999. Retrieval of
cloud microphysical parameters from 94-GHz radar
Doppler power spectra. J. Atmos. Oceanic Tech., 16, 489503.
Gossard E. E., 1994: Measurements of cloud droplet size
spectra by Doppler radar. J. Atmos. Oceanic Tech., 11,
712-726
2 = 2t + 2DSD.
Figure 13: Modal radius time series of the
drizzle modal radius (m) at gate 3, 260 m
over the time interval 13:20 UTC to 14:30
UTC.
Frisch, A.S., C.W. Fairall and J.B. Snider, 1995.
Measurement of stratus cloud and drizzle parameters in
ASTEX with a K –band Doppler radar and microwave
radiometer. J. Atmos. Sci., 52, 2788-2799.
Istok, M.J. and R.J. Doviak, 1986. Analysis of the relation
between doppler spectral width and thunderstorm
turbulence. J. Atmos. Sci., 42, 607-614.
Kollias, P. and B.A. Albrecht, 2000. The turbulent structure
in a continental stratocumulus cloud from millimeterwavelength radar observations. J. Atmos. Sci., 57, 24172434.
Fig. 11 Time series of the standard deviation of the mean Doppler velocity
(average time 5 minutes) (dashed line) and standard deviation of the Doppler
spectrum (solid). The circles at the top of the graph show the drizzling
periods.
Figure 12: Microphysical retrievals of
drizzle modal radius (m), number density
(lt-1), logarithmic width (x), and liquid
water content (gm-3). The data from the
radar at each gate are averaged over the
time interval 13:40 UTC to 13:45 UTC.
Fig. 14: Time series of Mean Doppler velocity
and Doppler spectrum skewness during a
drizzling period.
Kollias, P., B.A. Albrecht, R. Lhermitte and A. Savtchenko
2001. Radar Observations of Updrafts, Downdrafts, and
Turbulence in Fair Weather Cumuli. J. Atmos. Sci., 58,
1750-1766.
Randall, D. A., Q. Shao, and C-H Moeng, 1992: A second
order bulk boundary-layer model. J. Atmos. Sci., 49, 19031923.