Measuring Entrainment, Divergence, and Vorticity on the Mesoscale from Aircraft

Donald H. Lenschow National Center for Atmospheric Research, * Boulder, Colorado

Search for other papers by Donald H. Lenschow in
Current site
Google Scholar
PubMed
Close
,
Paul B. Krummel CSIRO Atmospheric Research, Aspendale, Victoria, Australia

Search for other papers by Paul B. Krummel in
Current site
Google Scholar
PubMed
Close
, and
Steven T. Siems Department of Mathematics, Monash University, Clayton, Victoria, Australia

Search for other papers by Steven T. Siems in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

Three independent techniques for measuring entrainment at the top of the planetary boundary layer (PBL) from an aircraft are discussed: 1) measuring the terms in the budget of a scalar and solving for the entrainment term; 2) estimating the entrainment velocity as the negative of the ratio of a scalar flux at the top of the PBL to the jump in its mean value across the top; and 3) measuring the divergence within closed (circular) flight paths, integrating with height to obtain the mean vertical motion at the PBL top, and estimating the time rate of change of the PBL top to solve for the entrainment velocity. All of these techniques can be implemented using the same flight pattern. The first two techniques have been used with some success previously, but the divergence technique, as far as the authors know, has not been used for entrainment measurements. The closed flight track can also be used to measure vorticity, with somewhat greater accuracy than for divergence, since the vorticity is typically several times larger than the divergence. These techniques were implemented using the National Center for Atmosperic Research C-130 in the Aerosol Characterization Experiment (ACE-1) and the results for the divergence technique are discussed. It is shown that measuring divergence and vorticity with an aircraft is feasible but is at the edge of currently used air motion sensing technology.

Corresponding author address: Dr. Donald H. Lenschow, NCAR, P.O. Box 3000, Boulder, CO 80307-3000.

Abstract

Three independent techniques for measuring entrainment at the top of the planetary boundary layer (PBL) from an aircraft are discussed: 1) measuring the terms in the budget of a scalar and solving for the entrainment term; 2) estimating the entrainment velocity as the negative of the ratio of a scalar flux at the top of the PBL to the jump in its mean value across the top; and 3) measuring the divergence within closed (circular) flight paths, integrating with height to obtain the mean vertical motion at the PBL top, and estimating the time rate of change of the PBL top to solve for the entrainment velocity. All of these techniques can be implemented using the same flight pattern. The first two techniques have been used with some success previously, but the divergence technique, as far as the authors know, has not been used for entrainment measurements. The closed flight track can also be used to measure vorticity, with somewhat greater accuracy than for divergence, since the vorticity is typically several times larger than the divergence. These techniques were implemented using the National Center for Atmosperic Research C-130 in the Aerosol Characterization Experiment (ACE-1) and the results for the divergence technique are discussed. It is shown that measuring divergence and vorticity with an aircraft is feasible but is at the edge of currently used air motion sensing technology.

Corresponding author address: Dr. Donald H. Lenschow, NCAR, P.O. Box 3000, Boulder, CO 80307-3000.

1. Introduction

Measurements of entrainment and mesoscale flow features from aircraft play an important role in furthering our understanding of the atmosphere. Aircraft have unique capabilities for obtaining measurements on the mesoscale in the lower troposphere over almost any geographic location. There is a long history of work on this topic, but recent technical developments have allowed us to consider new approaches to obtaining useful measurements for these applications. At the same time, new results from numerical modeling and tank experiments have pointed the way toward new observational requirements.

One of the important scientific areas for global climate research is clouds. A specific example is the important role that low-level marine stratus clouds play in the earth’s radiation budget. They increase the short-wave albedo compared to the underlying ocean but have little effect on the longwave radiation emitted to space. Satellite analyses have shown that these clouds can cause a change in net cloud forcing of about −100 W m−2 in areas affected by these clouds and contribute about −17 W m−2 to the global net cloud forcing (Klein and Hartmann 1993). Their climatic importance has led to several intensive field studies off the California coast [e.g., Dynamics and Chemistry of Marine Stratocumulus Experiment (DYCOMS) and First ISCCP (International Satellite Cloud Climatology Project) Regional Experiment (FIRE)] and in the vicinity of the Azores [the Atlantic Stratocumulus Transition Experiment (ASTEX)], which have given us further insight into processes that control their evolution and provided us with datasets for comparison with results from modeling and numerical simulation. It is difficult, however, to include the effects of these clouds in models since the clouds are often less than 500 m thick and are intimately connected to boundary layer processes.

At a workshop on modeling and large-eddy simulation of boundary layer clouds [National Center for Atmospheric Research (NCAR)–Global Energy and Water Cycle Experiment (GEWEX) Cloud System Study (GCSS) Workshop on Boundary Layer Clouds, co-chaired by C.-H. Moeng and W. R. Cotton, 16–18 August 1994, reported by Moeng et al. (1996)], most participants agreed that the single most important variable that needed observational confirmation was the mean entrainment velocity We at the top of the cloud layer [i.e., the rate at which the planetary boundary layer (PBL) accretes fluid from the overlying nonturbulent free atmosphere by turbulent mixing processes]. The entrainment velocity is crucial in predicting the evolution of stratiform cloud, and yet different models predict drastically different entrainment rates, with corresponding wide variations in cloud cover and liquid water content. This is a major obstacle to further progress in developing generalized formulations for application to large-scale models. In order to determine which of the various numerical schemes give realistic results, further observational work is required since past field studies did not adequately address this issue.

Other examples of where measurements of We are needed include the clear-air PBL where the entrainment rate may be enhanced by shear across the PBL top or by breaking waves, and trade wind and fair-weather cumulus regimes where the PBL dynamics is strongly affected by cloud dynamics. Russell et al. (1999) have shown that in some cases entrainment can occur in both directions. They documented the existence of an intermediate layer, which they called the buffer layer (BuL), lying between the PBL and the free troposphere. This layer was intermittently turbulent and they documented entrainment from the BuL to the PBL, as well as from the PBL to the BuL, and from the free troposphere to the BuL. Another application is estimating trace species budgets in the PBL. A significant term in these budgets is often the entrainment flux, which is the species flux across the top of the PBL and is equal to the entrainment velocity times the difference in concentration across the top of the PBL. An example of this is evaluation of the chemical sink of ozone in the marine stratus–capped PBL by Kawa and Pearson (1989b) and photochemical production of ozone in the clear-air PBL over land by Lenschow et al. (1981).

This paper addresses this need by proposing an observational strategy that can estimate We by three independent techniques—all of which can be implemented simultaneously on the same flight track with one aircraft. In addition, one of the techniques proposed here—the divergence technique—involves measurements of divergence and vorticity on the mesoscale, which can also be of intrinsic interest. Examples of where these measurements are important include the study of mesoscale phenomena such as deep cumulus convection, land/sea breeze circulations, orographically modified flows, mesoscale cellular convection, and cold-air outbreaks over the ocean. Schubert et al. (1979) have shown that model predictions of mixed-layer depth and cloud thickness of marine stratocumulus are very sensitive to the PBL divergence.

Measuring divergence from an aircraft has, in the past, been difficult to implement from aircraft because of the stringent accuracy requirements and has not been used for estimating entrainment velocity. Both instrumental and sampling errors contribute to this difficulty. Therefore, in the next section, after we review two of the techniques that have been used and described previously, we go into more detail on the divergence technique. Section 3 discusses observations of divergence and vorticity, obtained from the NCAR C-130 during the Aerosol Characterization Experiment (ACE-1), which were used to test the technique.

2. Measuring entrainment velocity

A major problem in estimating We from observations is the difficulty in estimating a velocity of such small magnitude. Typically, for marine stratus We < 0.01 m s−1. Kawa and Pearson (1989a), for example, obtained a mean entrainment velocity of 0.003 m s−1 for a set of eight flights from the Dynamics and Chemistry of Marine Stratocumulus Experiment (DYCOMS) off the California coast in summer. Although they obtained reasonable estimates of We for the set of flights, they estimated uncertainty for each case to be at least 50% of the mean. This was the result of both the measurement strategy, which was not optimized for this measurement, and their use of ozone and total water as tracers of entrainment.

Another constraint for this measurement is that all the techniques discussed here are best carried out in a horizontally homogeneous region well away from complicated terrain; for example, over the ocean well away from shoreline or island effects, and in regions away from fronts, in order to obtain reasonably stationary conditions. This points toward a mobile measurement capability, such as from an aircraft or ship. Major disadvantages of a ship are 1) the difficulty in obtaining in situ measurements near and above the PBL top and 2) the time required to obtain statistically significant measurements (although a ship-mounted lidar or radar using a circular scan at a constant elevation angle might be useful for divergence measurements). Therefore, we consider here an airborne deployment.

It is important in making a measurement of this type that we have confidence in the answer. This is particularly true in a situation such as this, where there is no history of being able to make this measurement with sufficient accuracy to make comparisons with, for example, model predictions or synoptic analyses. One good way of achieving this is to have several independent techniques for obtaining the answer. If they agree, this gives us some assurance that the measurement is correct. Fortunately, there are three independent techniques, described in the next three sections, that can be implemented simultaneously using the same flight track from one aircraft for this measurement.

a. Budget technique

The first is the budget technique. In this approach, the entrainment rate is obtained by estimating all the sources and sinks of a tracer species, S(x, y, z, t) = S(z) + s(x, y, z, t), where S is the mean concentration at height z in the PBL and s a fluctuation in the mean. One of the terms is the entrainment flux (ws)zi, where the overbar denotes an average over an area (or along a line) that is large (or long) enough to give a statistically significant value. As shown by Russell et al. (1999), assuming a discrete jump over an infinitesimal layer at the top of the PBL, δS, the entrainment velocity at the top of the PBL is
i1520-0426-16-10-1384-e1
where δS = Szi+ − {S}, Szi+ is the concentration just above the PBL and Qs is the chemical source or sink of S. The curly braces { } denote an average throughout the PBL over the time period of measurement and zi is the average depth of the PBL over the period of measurement. Russell et al. (1999) extended this approach to a two-layer system in which entrainment occurs in both directions across the top of the PBL, and the second layer, which is intermittently turbulent, extends up to a second interface, across which entrainment occurs only from the free atmosphere into the second layer. Boers et al. (1999) and Bretherton et al. (1995) also discuss applications and limitations of this technique. The other two techniques are described in the next two sections. Further details on implementing these two techniques are given by Lenschow (1996).

b. Flux/δS approach

The second is a direct measurement technique where a conserved (on a timescale of several hours or more) tracer is used to obtain flux profiles through a well-mixed PBL. The entrainment velocity is calculated from the ratio of the flux extrapolated to the top of the PBL (ws)zi to the mean difference in tracer concentration across the top (Lilly 1968):
i1520-0426-16-10-1384-e2
This is the approach used by Kawa and Pearson (1989a). Comparing their two estimates for each flight (one using total water qt and the other ozone as tracers), we see that in most cases they are in reasonable agreement—the average difference is 0.0007 m s−1. However, both ozone and qt have problems for this application. They both have significant concentrations in the overlying free atmosphere, which tend to vary significantly both horizontally and vertically because of the absence of mixing between decoupled layers that have diverse origins and the intermittency of events that produce vertical transport. Ozone in particular is generated mostly in the stratosphere and removed mostly at or near the earth’s surface so that the concentration just above the PBL is typically greater than within the PBL (e.g., Kawa and Pearson 1989b). The result is that the jump across the top of the PBL typically shows considerable horizontal variability. In addition, it has only a small sink at the surface [a deposition velocity, wd ≡ −(wO3)0/O3 ≃ 3 × 10−4 m s−1, where (wO3)0 is the surface flux of ozone and O3 is the mean concentration at some reference height in the surface layer; Kawa and Pearson (1989b)]. Another problem with using qt is loss of liquid water from the cloud layer through drizzle.

The ideal characteristics of a tracer species would be 1) a horizontally homogeneous surface source; 2) a lifetime of 104 < τ < 106 s, short enough to ensure a large jump across the PBL top and minimize horizontal variations in concentration above the PBL and yet with a long lifetime as compared to the turbulence mixing time in the PBL zi/w∗, where w∗ is the convective velocity scale; and 3) negligibly soluble in clouds. Two examples of naturally occurring scalars that have these characteristics are CH3SCH3 (dimethyl sulfide; DMS) and CH3I (methyl iodide). In addition, of course, it is necessary to be able to measure fluxes of these species—preferably via the eddy-correlation technique (Lenschow 1995), which requires fast-response (≥5 Hz) sensors. Although eddy-correlation fluxes of DMS have not yet been measured directly, Hills et al. (1998) have developed a prototype fluorine-induced chemiluminescence DMS sensor that has sufficient sensitivity and time response for eddy flux measurement. In the absence of direct flux measurements, we do not have good spatial definition of DMS emission; however, it is closely related to oceanic concentration of DMS, which has been found to be reasonably uniform on a scale of hundreds of kilometers (e.g., Bates et al. 1987). No technique is currently available to measure eddy flux of CH3I.

Straightforward eddy correlation is not the only possibility for measuring the entrainment flux. Other techniques exist that do not require such fast-response measurements (Lenschow 1995).

  • Conditional sampling—collecting air samples in more than one reservoir according to some property of the vertical velocity. Eddy accumulation denotes collecting samples at a rate proportional to the instantaneous vertical velocity in two reservoirs—one for positive w and the other for negative w. The flux is proportional to the difference in concentration between the two reservoirs. A “relaxed” version of eddy accumulation collects samples at a constant rate in either of two reservoirs according to whether w is positive or negative.

  • Intermittent or disjunct sampling—grabbing a sample quickly, then measuring the tracer concentration more leisurely. High-frequency response is maintained by the fast sampling. If the measurement is made once per integral scale, which is typically about 1 s for aircraft measurements, the random error due to finite sample length is not appreciably increased (Lenschow et al. 1994). Cooper and Shertz (1995) describe an airborne system that can be used for both eddy accumulation and intermittent sampling.

  • Mixed-layer gradient or variance—using mean concentration measurements from a minimum of three levels through the mixed layer or variance measurements from a minimum of two levels to solve for the surface and entrainment fluxes using the top–down—bottom–up (TD–BU) gradient or variance relations, respectively, first proposed by Wyngaard and Brost (1984), and later refined by Moeng and Wyngaard (1984, 1989) using large-eddy numerical simulations (LES). The gradient relations can be integrated to solve for the surface and entrainment fluxes in terms of the two concentration difference measurements (Davis 1992; Davis et al. 1994), while the variance relations can be applied directly to the measured variances to obtain the entrainment velocity (e.g., Davis et al. 1997). The variance technique technique is, however, sensitive to possible mesoscale contributions to the variance, and hence to the long-wavelength cutoff used in estimating the variance.

1) DMS as a tracer

Examples of the use of DMS mean profiles are presented by Russell et al. (1999) and Lenschow et al. (1999). Because of the ∼1–2-day (105 to 2 × 105 s) lifetime of DMS, air above the PBL is likely to have negligible concentration of DMS. This is particularly true of the air overlying the stratus regions associated with the summertime subtropical anticyclone over the major ocean basins since the stably stratified subsiding flow has likely not been in contact with PBL air for many days previously. In this case, the budget equation for a chemically reactive species in a horizontally homogenous PBL reduces to
i1520-0426-16-10-1384-e3
where τ is the species decay time. Integrating from the surface to the mean top of the PBL zi, and assuming steady-state conditions,
i1520-0426-16-10-1384-e4
Equation (4) can be solved for the flux at the top, (ws)zi = −WeδS. For DMS, we assume that Szi+ = 0, so that δS ≃ −S. Therefore,
i1520-0426-16-10-1384-e5
For (ws)0 = 2 pptv m s−1, (≃6 μmole m−2 day−1), We = 0.005 m s−1, τ = 2 days, and zi = 103 m, we obtain from (5) S ≃ 185 pptv; and from (4), (ws)zi ≃ 0.93 pptv m s−1. That is, the flux at zi is typically about half the surface flux, which means that the entrainment velocity should be relatively well defined and easy to estimate.
We also note that the budget Eq. (3) can be integrated between two levels Δz = z2z1 and solved for the reactivity time constant τ,
i1520-0426-16-10-1384-e6
Kawa and Pearson (1989b) used this technique to estimate the lifetime of ozone (10–32 days) in the PBL during DYCOMS.

Finally, DMS has great potential as a tracer for addressing other problems. For example, deep convection provides a conduit for transporting PBL air to the upper troposphere on a timescale of an hour. Measuring DMS in the outflow regions of such convective events over the ocean could provide a means for estimating the amount of PBL air brought up to these levels. Another example is its use as a tracer for measuring TD–BU functions in the atmosphere. Thus far, these functions have been estimated from LES and laboratory convection tank experiments (Piper et al. 1995), but not in the atmosphere. Combining DMS with other species that have quite different behavior in the PBL provides the necessary contrast in structure that both top–down and bottom–up functions can be solved for simultaneously. Ozone provides a good contrast since it has a surface sink and typically a large negative entrainment flux.

2) Estimate of the sampling error

The error variance for a scalar flux measured over a finite sample length L is given by (Lenschow and Kristensen 1985)
i1520-0426-16-10-1384-e7
where ℓw is the integral length scale of the vertical velocity, and σ2w and σ2s are the variances of w and a scalar s. Lenschow and Stankov (1986) found empirically that
i1520-0426-16-10-1384-e8
where z∗ = z/zi. We can estimate σs from the TD–BU formulation obtained by Moeng and Wyngaard (1984, 1989) using large-eddy numerical simulations that relate mixed-layer variance of a scalar to surface and entrainment fluxes,
i1520-0426-16-10-1384-e9
where fbz−0.9*, ft ≃ 3.1(1 − z∗)−3/2, and fbt ≃ 1.5 are the dimensionless bottom–up and top–down variance functions, and the TD–BU covariance function, respectively. Solving for the normalized scalar variance we obtain
i1520-0426-16-10-1384-e10
where α = (ws)0/(ws)zi.

This normalized variance is plotted in Fig. 1. Using the above case as an example, α = 2.2, and we let the convective velocity scale w∗ = 1 m s−1 and z∗ = 0.5. The result is σ2sw2*/(ws)2zi ≃ 24. This gives σs ≃ 4.5 pptv. The relative error for a 200-km flight path with σw = 0.7 m s−1 is σF/(ws)zi ⩽ 0.20. Thus, the random error variance due to a finite sample length for flux measurement of DMS can be made small enough that it is possible to obtain estimates of We with reasonable accuracy.

c. Divergence technique

The third approach is measuring the mean vertical velocity at the top of the PBL Wzi and the mean time rate of change in PBL height dzi/dt. The entrainment velocity is
WedzidtWzi
The change in PBL height can be measured either in situ by aircraft penetrations through the capping inversion or remotely with, for example, lidar throughout the course of a research flight. Typically, the magnitude of Wzi is similar to that of We. There is no current or planned measurement technique that has sufficient accuracy for measuring Wzi directly. However, the vertical velocity at cloud top can also be estimated, in principle, by integrating the horizontal divergence through the PBL.
From the continuity equation for an incompressible fluid,
i1520-0426-16-10-1384-e12
where A is the area enclosed by the integration path, is the area-averaged vertical velocity within the integration path, and υn is the horizontal velocity component normal to the path of integration. The most efficient aircraft flight path for this measurement is a circle since 1) a circle has the largest enclosed area of any closed curve of a given circumference (e.g., 27% more area than a square of the same perimeter) and 2) no time need be wasted in turns since the rate of turning is slow enough that there is no reduction in accuracy as may occur during a sharp turn. For a circular flight path of radius R,
i1520-0426-16-10-1384-e13
where Vn is the mean normal wind component averaged around the closed flight track. For constant divergence from the surface to height z, the mean vertical velocity at z is
i1520-0426-16-10-1384-e14

This vertical velocity estimate assumes that the horizontal velocity field is constant with time in the coordinate system of the measurement throughout the measurement period. We can most closely realize this assumption by flying the closed path in a Lagrangian framework—that is, flying the aircraft in a circular pattern where the center of the circle drifts with the mean wind.

Assuming a steady-state PBL with a horizontally homogeneous divergence field, the area averaging used in (14) should give the same results as a linear averaging process (e.g., a direct measurement of vertical velocity) for a large area as compared to the square of the horizontal length scale of w fluctuations and a flight path much longer than the length scale of fluctuations. Therefore, we assume that at the top of the PBL wzif = Wzi. Typical values for −Wzi are of order 0.01 m s−1. Ideally we would like to be able to measure to within 20% of this, or ∼0.002 m s−1. Thus, for z = zi = 103 m and R = 180/2π ≃ 28.7 km (which corresponds to a half-hour flight duration for an airspeed of 100 m s−1), this translates to an accuracy in Vn of R/2zi times the accuracy in Wzi. Since the magnitude of the normal velocity component is proportional to the ratio R/zi, the larger the circle, the larger the acceptable error in the measurement of Vn. In the above example, Vn = 15 × Wzi, so the acceptable error in Vn is
εVn−1

1) Sampling errors

Two approaches to measuring the divergence through the PBL are to 1) fly a series of closed (circular) flight paths at several levels in the PBL and 2) obtain vertical soundings of the normal wind component at various points on the periphery of the area under consideration, for example, from dropwindsondes launched from an aircraft flying above the PBL or rawindsondes launched from a series of stations surrounding the area. The error variance of a variable υ (here we drop the subscript from υn, but have in mind still the horizontal velocity component normal to the flight path) is defined as
σ2υ, LυLυ2
where L is the sample length and 〈 〉 denotes an ensemble average. For an ergodic variable measured by, for example, an aircraft, (16) can be reduced, for L ≫ ℓυx, to (Lenschow et al. 1994)
i1520-0426-16-10-1384-e17
where ℓυx is the integral length scale of υ along the flight path and σ2υ is the variance of υ. The integral scale along the flight path is defined as
i1520-0426-16-10-1384-e18
where ρυx(ξ) is the autocorrelation function, which for L ≫ ℓυx, can be written as
i1520-0426-16-10-1384-e19
Lenschow and Stankov (1986) found that the integral scale for horizontal velocity components measured in the horizontal plane is
υxzi
For typical marine stratocumulus, we assume σ2υ ≃ 1 m2 s−2 and zi = 1000 m. For a circular flight path of radius R = 29 km (at an airplane speed of 100 m s−1, this takes one half-hour per revolution),
σ−1
We also note that eight revolutions would reduce this number to 0.025 m s−1. Thus the standard deviation due to sampling length for eight revolutions is less than the value specified in (15).
The second approach, using soundings, suffers from large sampling errors unless an inordinate number of soundings is obtained. If the distance between soundings Δ ≫ ℓυx, then from the results of Lenschow et al. (1994), assuming vertical homogeneity the random error variance based on a set of instantaneous independent measurements over the horizontal distance L becomes
i1520-0426-16-10-1384-e22
Each sounding, however, may give more than one independent estimate of υn as it passes through the PBL. We can incorporate this by modifying (22) to read
i1520-0426-16-10-1384-e23
where n is the number of independent estimates of υn obtained per sounding.
We can estimate the sampling error of a set of soundings in the following way. The autocorrelation function between horizontal velocity components along the vertical direction, again assuming ergodicity, can be written as
i1520-0426-16-10-1384-e24
Admittedly, the assumption of ergodicity is less tenable in the vertical than in the horizontal since the velocity field undergoes changes both near the surface and near the top of the PBL. However, observations (e.g., Lenschow et al. 1980) show that through the bulk of the mixed layer, statistics of the horizontal velocity components vary only slightly with height.
The autocorrelation function can also be expressed in terms of the Fourier transform of the coherence function coh(k, ζ) and the spectrum of υn along the aircraft track, given by Φυ(k) (Panofsky and Dutton 1984). If we assume that the phase angle between the horizontal velocity components at different levels in the PBL is negligible (i.e., that the turbulent eddy structures are not tilted in the vertical),
i1520-0426-16-10-1384-e25
We assume a simple power spectrum of the form (Mann and Lenschow 1994)
i1520-0426-16-10-1384-e26
Lenschow and Kristensen (1988) proposed the following formulation for the coherence function:
i1520-0426-16-10-1384-e27
where the constant b = 0.7468. Davis (1992) estimated the coherence of w from two vertically stacked aircraft flying in formation (the experiment is described by Lenschow and Kristensen 1988). He found that for az ≃ 0.25, Eq. (27) is a reasonable fit to the coherence of w along the vertical axis, using the integral scale for w in the horizontal plane obtained by Lenschow and Stankov (1986), which is given by (8). In the middle of the mixed layer, ℓw/zi ≃ 0.17. We assume that we can replace ℓυ in (27) with the midmixed-layer value of ℓw. This would be strictly true if the turbulence were isotropic; even for vertically “squashed” turbulence it is probably true within a factor of about 2 (Kristensen et al. 1983). Here we are interested only in an approximate estimate of the correlation function, so substituting ℓw/zi = 0.17 along with (26) and (27) into (25) and numerically integrating, we obtain the autocorrelation function in terms of the normalized separation distance ζ/zi, which is plotted in Fig. 2.

Integrating the vertical autocorrelation function (25) over the normalized separation distance, the integral scale for υ in the vertical direction is ℓυz/zi ≃ 0.34, about 75% of ℓυx/zi. Thus, based on this very approximate analysis there is no evidence for any strong asymmetry in the horizontal velocity structure through the mixed layer. The result indicates that in principle it is possible to obtain about three statistically independent estimates of υ per sounding through the PBL. [This also means that three aircraft flying coincident closed flight paths simultaneously (one near the surface, one in the middle, and the other near the top of the PBL) will make nearly independent measurements of υn, and that remote measurements of a vertical cross section of the normal component of the horizontal wind by, e.g., an airborne Doppler lidar or radar will approximately triple the sample size of an in situ measurement.]

Equating the error variance from continuous sampling (17) to that from instantaneous measurements (23) with the number of independent samples per sounding, n = 3, we obtain Δ ≃ 6ℓυx. That is, a sounding is required about every 3 km for a 1-km-deep PBL to match the sampling error from a continuous horizontal velocity measurement.

From (11), to measure We, we also need to measure dzi/dt. For a 4-h period, to measure We to 0.001 m s−1, the change in zi must be measured to Δzi ≃ 15 m. This accuracy is possible to achieve using either a downward-pointing lidar on a traverse over the area of interest or sufficient penetrations through the top of the cloud layer. We can estimate the requisite number of penetrations in the following way. Lenschow (1990) reports the standard deviation of cloud-top height measured by lidar for two cases in the California stratocumulus to be σz = 13 and 26 m. If we assume conservatively that σz ⩽ 30 m, then if zi is normally distributed, the error for N independent estimates of zi is σz/N. Thus, in general, a minimum of about four independent penetrations through the cloud top should be sufficient for estimating zi to 15-m accuracy for California-type stratocumulus. In certain situations, however, estimating the change in PBL height may become more problematic since for σz = 60 m, the number of penetrations increases to 16.

2) Vorticity

The same technique discussed above can be used for measuring the circulation or the vertical component of vorticity within the closed flight path by substituting the velocity component parallel to the flight track υt (positive for counterclockwise rotation) for υn in (11),
i1520-0426-16-10-1384-e28
This may be useful also as an independent estimate of divergence, since the equation for the vertical component of absolute vorticity, neglecting the tilting and solenoidal terms and terms involving the horizontal momentum transport by turbulence, can be written as (e.g., Holton 1972)
i1520-0426-16-10-1384-e29
where f is the Coriolis parameter. Normally, fζυ. Therefore, if we assume steady state and a linear decrease in the stress term with height to zero at zi, integrate vertically, and approximate the stress term by a drag formulation, we obtain
i1520-0426-16-10-1384-e30
where CD is the drag coefficient and Um is the mean horizontal wind speed at the height for which the drag coefficient applies.
Ching (1975) applied this approach to measurements of rawinsonde wind profiles from ships obtained during the Barbados Oceanographic and Meteorological Experiment to estimate the drag coefficient. Here we point out that for the idealized situation indicated above, the divergence is approximately related to the vorticity by the relation
i1520-0426-16-10-1384-e31
Strong divergence occurs with anticyclonic circulation;in the Northern Hemisphere, positive divergence implies negative vorticity. For typical values in a nonequatorial PBL, (31) indicates that the magnitude of the vorticity is typically from several times to an order of magnitude larger than divergence and thus can be measured more accurately.

3) Instrumental errors

We now consider errors in the mean divergence measurement from the perspective of likely errors in the airborne sensors. The actual horizontal wind measured by an aircraft is obtained from the difference between the velocity of the airplane in a geographic reference frame [measured, e.g., by an inertial navigation system (INS) with long-term corrections provided by the Global Positioning System (GPS)] and the velocity of the air in the airplane coordinate system (obtained, e.g., from differential pressure measurements between ports on a nose radome). Here we are interested only in the component normal to the airplane flight track υn. Errors in the airplane velocity measurement (from INS updated with GPS) tend to be independent of aircraft heading. Therefore, in considering their impact, we consider the simple case of a horizontal velocity field that is constant in direction but increases linearly in the x direction, U = U0 + U1(x/R), as shown in Fig. 3. (Since the flight path is assumed to be circular, the wind direction we choose is arbitrary.) The normal wind component in polar coordinates is given by
i1520-0426-16-10-1384-e32
The divergence for this case is
i1520-0426-16-10-1384-e34
Assuming constant divergence through the PBL and a required measurement accuracy of Wzi = 0.001 m s−1, for R = 29 km (30-min revolution), and zi = 1000 m, the required accuracy is U1 = 0.03 m s−1, which is equivalent to ∼54 m per revolution. Currently available GPSs can measure to about 100 m absolute accuracy, which therefore seems marginal for estimating the horizontal airplane velocity averaged over a circle with R = 29 km. We point out, however, that higher-accuracy GPS information (which GPS is inherently capable of providing with the approval of the controlling governmental agency) would then be essential.
The next step is to consider errors in angle measurement. Probably the simplest error to consider is an error in the angle measurement of the normal wind component of a constant value ε, which might be due to, for example, a misalignment between the INS and the wind-sensing instrumentation, which gives the measured divergence
i1520-0426-16-10-1384-e35
Thus, for a relative error of ⩽10%, ε ⩽ 26° per revolution, so that an error from this source is not significant.
However, if we introduce a time-dependent error, or drift in the angular misalignment of ε(t) = αθ, then for small α, it can be shown that, by setting sinαθαθ and cosαθ ≃ 1,
i1520-0426-16-10-1384-e36
We let U0 = 5 m s−1. Then for a relative error of ⩽10%, ε ⩽ 0.33° per revolution. This is a much more significant source of error and is comparable to the accuracy of the INS measurement of true heading, which is the most likely source of such a drift.
Another consideration with a circular flight path is the magnitude of the roll angle ϕ that results from a constant-rate turn. This can be easily estimated by equating the centrifugal force mU2a/R, where Ua is the true airspeed and m the mass of the aircraft, to the component of the lift force that is opposite to the centrifugal force, mg tanϕ, where g is the gravitational acceleration. Solving for the roll angle,
i1520-0426-16-10-1384-e37
For Ua = 100 m s−1 and R = 29 km (equivalent to a 30-min flight time around the circle), ϕ ≃ 2°, which would have negligible effect on the accuracy of the air velocity measurement.

Finally, we consider errors in the measurement of the velocity of the air with respect to the aircraft. For divergence measurement, this is very nearly the horizontal component of the air velocity normal to the longitudinal axis of the aircraft (or the lateral component) since the difference between the aircraft heading and track angle is small. Currently, a standard technique for obtaining this measurement is to use pressure difference measurements from ports on a radome mounted on the nose of the aircraft to calculate the flow angle, called sideslip, β. The normal air velocity component with respect to the aircraft is then ≃βUa. Certainly the absolute accuracy of this measurement is greater than 0.006 m s−1, primarily since the system is calibrated by flight maneuvers, which probably cannot be used to achieve absolute accuracies better than a few tenths of a meter per second because of spatial and temporal variability in the wind field. However, the effect of a mean offset in β can probably be mostly cancelled out by flying circles in opposite directions. The remaining error is difficult to quantify and remains a question. A similar result holds for the effect of the true airspeed, which is also obtained from a differential pressure measurement (corrected for air density), on the longitudinal component of air velocity, and thus on the vorticity measurement.

An alternative approach for air motion measurement is a side-looking Doppler laser system, sometimes called a velocimeter. A forward-looking laser velocimeter was successfully flown on the NCAR Sabreliner (Keeler et al. 1987), and design criteria for a more complex scanning laser velocimeter are discussed by Kristensen and Lenschow (1987). The advantages of this approach are 1) the measurement volume can be displaced several meters away from the aircraft where the flow is not perturbed by the aircraft and thus flight maneuvers are not required to calibrate the sensor, 2) Doppler shift is inherently an absolute measure of velocity with no hysteresis or environmental sensitivities, and 3) cloud droplets should not affect the accuracy. This is an attractive option that offers the potential of taking full advantage of the recent improvement in measurement accuracy of the airplane velocity through GPS. Currently, air motion measurement accuracy is the limiting factor in wind measurement from aircraft. Improved wind measurement would have application to many other research problems as well, such as measuring mesoscale wind fields, aerodynamic transfer coefficients, and mean advection.

d. Optimal flight track

Summarizing the results of the previous sections into a reasonable scenario for an observational study, the requirements for an optimal flight pattern to measure entrainment velocity in a marine stratocumulus regime are as follows.

  • The three independent techniques for estimating We should be simultaneously implemented with a single aircraft flight pattern.

  • The flight pattern should be as close to optimal as possible for all the techniques.

  • The flight pattern should permit correlative measurements essential to a complete documentation of the processes involved in generating entrainment.

One flight pattern that fits these criteria is a set of circles, flown in a Lagrangian frame of reference, of 30- to 150-km diameter. A convenient method to fly such a pattern is to put the airplane into a constant-rate turn (e.g., 12° per minute for a 30-min turn). A minimum of four circles spaced approximately uniformly through the PBL would allow two sets of two circles flown in opposite directions so that the biases in the longitudinal and lateral velocity components could be removed. Another set of circles flown above the PBL, one at the beginning and the other at the end of the PBL circles, would be used for estimating dzi/dt by a nadir-pointing lidar. Embedded within these circles would be occasional dips through the inversion to document the jump in scalars and velocity across the top of the PBL. If lidar is not available, the dips can also be used for estimating dzi/dt. If time is not available, flight paths bisecting the circle could be substituted.

3. Observations

a. ACE-1 case study

The Aerosol Characterization Experiment (ACE-1), carried out primarily over the ocean south of Australia during November–December 1995 (Bates et al. 1999), afforded an opportunity to test the techniques proposed here for measuring entrainment, divergence, and vorticity. One objective of ACE-1 was to estimate budgets of trace atmospheric constituents in the PBL over time periods of a day or more using the NCAR C-130 aircraft. This required a series of flights in the same air mass whose trajectory was labeled by means of constant-level balloons released from a ship. These sets of flights are called Lagrangian experiments. Strictly speaking, the trajectory is not truly Lagrangian, since the balloons do not stay in the same air mass, but rather stay at approximately the same altitude. However, because of efficient mixing, the PBL normally has a fairly uniform velocity, and the trajectory should be close to the trajectory of the mean PBL flow. An important term in the constituent PBL budgets is the contribution of entrainment through the top of the PBL, for which the circular flight patterns, centered on a balloon’s position as determined by GPS, were employed. Russell et al. (1999) have already used the C-130 measurements during the second set of Lagrangian measurements (flights 24–26) to compare entrainment rates calculated by all three techniques, and Wang et al. (1999a) have compared the flux and divergence techniques for the first Lagrangian period (flights 18–20).

To illustrate the calculation of divergence and vorticity for these circular flight patterns, we focus on flight 18 of ACE-1, the first flight of the first Lagrangian experiment. The NCAR C-130 flew southwest of Tasmania with the first circle centered near 44.8°S, 143.5°E. Over a 6-h period on 1 December 1995, the C-130 completed two sets of circles while drifting with the ambient wind so that the C-130 sampled nearly the same air mass throughout the flight. Each set consisted of five circles at various levels in the lower troposphere, with one in the free troposphere, in a quasi-Lagrangian frame of reference as estimated from an approximately constant-level balloon. The pressure altitude and flight track are shown in Fig. 4, with the circle number indicated on the pressure altitude plot.

A detailed account of the dynamics and thermodynamics of the PBL for flight 18 can be found in Wang et al. (1999a) hereafter W99a; Wang et al. 1999b. To summarize, the Lagrangian began in an air mass behind a weakening front. The flight logs note a complex PBL structure with decoupling occurring at the top of a surface-based turbulently mixed layer and two layers of cloud (broken stratocumulus and scattered cumulus). A change in horizontal wind of over 10 m s−1 was observed across the lower troposphere below the free atmosphere. Details of the synoptic meteorology and its impact on the Lagrangian trajectory can be found in Businger et al. (1999) and Siems et al. (1999; hereafter S99).

The set of divergence and vorticity estimates obtained for each circle is shown in Figs. 5a and 5b, where the solid dots indicate clockwise and the squares counterclockwise circles. Here, and in subsequent figures and tables showing ACE-1 data, we have reversed the sign of the vorticity so that the results conform to what would be expected in the Northern Hemisphere. Numbers next to the points indicate the circle number. We calculated means and standard deviations of the divergence and vorticity for flight 18, excluding circles 5 and 10, which were above the PBL. The standard deviations (the values below preceded by ±) were obtained by fitting a least squares linear fit separately to the set of four clockwise and counterclockwise divergence and vorticity measurements versus static pressure, then calculating the standard deviations of the remaining fluctuations. These standard deviations were then used to calculate the standard deviations of a set of eight estimates, using the assumption that the ratio of the sampled mean to the sampled variance follows a Student t-distribution (e.g., Bendat and Piersol 1971).

The resulting divergence is 5.1 ± 5.4 × 10−6 s−1. In this case, even the sign of the mean value has little significance; however, it is consistent with our expectation of a positive divergence in a postfrontal environment as the cold air from higher latitudes subsides behind the front. The average vorticity for the same set of circles is 1.9 ± 0.7 × 10−5 s−1 (Fig. 5b). In this case, the sign is significant; a clockwise (cyclonic) rotation, which is also consistent for airflow just behind a front in the Southern Hemisphere, but in disagreement with Eq. (31), which predicts counterclockwise rotation, likely because the terms neglected in obtaining (31)—that is, the time rate-of-change, tilting, and solenoidal terms—are significant in a situation such as this close to a front.

From (13), the normal wind component averaged around the closed flight track calculated from the estimated divergence error is Vn ≃ ±0.075 m s−1. This is about three times the theoretical sampling error estimated earlier for a set of eight circles. In the next section, we examine some reasons for the measured error being larger than predicted for this case.

b. Sensitivity to circle closure and mesoscale variability

As discussed earlier, neglecting instrument errors, the C-130 measurements should be sufficiently accurate to estimate the divergence and vorticity. There are, however, some further aspects specific to this case that need to be considered. First, we consider a particular circle from flight 18. Figure 6 shows the first circle from this flight plotted with the trajectory corrected for the mean wind measured by the aircraft. It would be highly unlikely for the circle to be precisely closed; however, the integrals in Eqs. (28) and (12) are still valid provided we can reasonably estimate a “closure point” near the end of the circle that minimizes any extraneous contribution to the closed integrals from overlap between the beginning and the end of the circle. Furthermore, (28) and (12) are valid for any closed curve. The criterion we used here for “closing” the circle was to integrate around the circle to the point of minimum distance from the starting point. The shortest distance we achieved was 13 m; in a few extreme cases during ACE-1 circles had gaps between starting and ending of over 1000 m. Figure 7 shows the running calculation of the divergence and vorticity for our example, which was within 175 m, or 0.1% of the total circumference, of closure.

To test the sensitivity of the calculations to the error in closure, we consider small changes to the length of the flight track. Table 1 shows divergences and vorticities as a function of time for this circle. At an aircraft speed of 100 m s−1, a 10-s error leads to a distance error of 1 km. Table 1 illustrates that for this case, the relative change in divergence with closure distance is larger than for vorticity, even though the actual change in vorticity is larger. This is a reflection of the order of magnitude larger amplitude of the vorticity. In this particular case, Fig. 8, which zooms in on the last few minutes of Fig. 7, shows that the normal wind component passes through zero at about 1744 UTC, which minimizes the closure error. We also see that divergence and vorticity are smoothly varying functions of time, so that it is feasible to estimate circle closure with considerable accuracy up to a few hundred meters. For this example, if the circle is closed to within about 3 s, or 300 m, the contribution of nonclosure to the measurement error of divergence and vorticity is negligible.

We note that the divergence sensitivity can be minimized by orienting the aircraft along the wind direction at the beginning of the circle so that there will be little contribution to the flux divergence. At the same time, this maximizes the component along the flight path so that the vorticity will be maximally sensitive to the circle closure. The results indicate that adding a few seconds flight time at the end to overlap the starting point is useful.

In theory, we could have also calculated the divergence (and vorticity) from semicircles that begin and end with the lateral (longitudinal) winds perpendicular to the airplane heading. Beginning and end points of the semicircle calculations, however, are hard to estimate. Combining this with the increased sampling error means that we would have little confidence in the results.

Figures 7 and 8 also give some indication of how sensitive the measurements are to mesoscale variability in the wind field. The winds in flight 18 are characteristic of a relatively turbulent marine PBL containing mesoscale variations. In Fig. 9, the autocorrelation functions and the integral of the autocorrelation functions are shown for the circle from flight 18 shown in Fig. 7. Before calculating the autocorrelations, the zero mode plus the first three harmonics of the Fourier transforms of the time series are removed from each component to eliminate modulation of the time series by the circular flight path. An inverse transform is then performed on the Fourier spectra to obtain the autocorrelation functions. The maxima in the integrals are estimates of the integral scales of both the normal (lateral) and the along-flightpath (longitudinal) wind components (Lenschow and Stankov 1986). We see that the values of the maxima are 13 (∼1.3 km) and 29 s (∼2.9 km), respectively. We note that these values are a small fraction of the distance around the circle, so that the effect of aircraft turning should be minimal. The ratio is 0.46, which is not far off from their predicted ratio in isotropic turbulence of 0.5 (e.g., Batchelor 1953)—of course, this does not imply that the turbulence is isotropic.

For this case, W99b estimated the depth of the PBL at about 900 m. Lenschow and Stankov (1986) used the average of the longitudinal and lateral integral scales in their analysis, and obtained ℓυx/zi = 0.45. Here we obtain ℓυx/zi ∼ 2.3, which is about five times what they obtained. We suspect that in this case there is significant mesoscale variability, which can sometimes occur in the horizontal wind (e.g., if the C-130 flew through or under a patch of cumulus clouds, or across an ocean temperature discontinuity), which means that the sampling error in this example is larger than that predicted by (17). Lenschow and Stankov (1986) obtained their data from a more unstable PBL, which consequently likely had less mesoscale variability. The boundary layer in flights 13 and 19 (Table 2), for example, had such large mesoscale variability that we have little confidence in their estimated divergence and vorticity. Thus far we have not been able to determine the extent to which systematic errors in the air velocity measurements contribute to the observed error. This could be accomplished by carrying out a more focused flight program, as described in our conclusions.

The estimates for flight 18 are an example of a consistent set of measurements. Often the estimates from the ACE-1 flights are not as easily interpreted, both because of more complicated synoptic situations and because the flight patterns were not optimal for application of this technique. Since the measurements are generally not sufficiently accurate to easily detect changes with height, we have not looked into what measurement accuracy might be possible for changes with height. In this case, however, the vorticity shows a consistent increase with height of about 4 × 10−8 m−1 s−1, which seems significant. The difference between the clockwise and counterclockwise circles is due to biases in the differential pressure measurement for sideslip (divergence) and the Pitot-static pressure measurement for airspeed (vorticity), and illustrates the importance of flying in both directions.

Generally, the vorticity profiles showed considerably less relative scatter than the divergence, which is consistent with the vorticity being typically as much as an order of magnitude larger than the divergence (31). It is also possible that the assumption of constant divergence across the PBL is invalid. Divergences produced from numerical models often have gradients across the lower troposphere (e.g. W99a). If the divergence does vary across the PBL, this could have serious implications for many numerical studies of the PBL. These models are known to be highly sensitive to the prescribed divergence (Schubert et al. 1979). We note that an appropriately designed experiment using circular flight tracks may be able to address whether there is a random or systematic variability in the divergence and vorticity in the lower troposphere.

c. Summary of ACE-1 measurements

Since ACE-1 was the first field experiment to extensively use circular flight paths for measuring divergence and vorticity in the PBL, we summarize the results in Table 2. Only flights with four or more PBL circles are included, and a few flights were rejected because the set of circles had problems (e.g., one clockwise and three counter-clockwise circles, or circles not in a Lagrangian framework). Circles from the free troposphere were not included.

Ideally, we would like to compare the C-130 measurements with an independent measure of the divergence and vorticity taken on the same scale. Unfortunately, that is not possible because of the remote location and the Lagrangian approach to the measurements. Our alternative is to compare with synoptic-scale calculations based on the global analyses used to initialize numerical weather prediction (NWP) models. These are the values often used in numerical models of the PBL. From the horizontal wind analyses, it is straightforward to derive the divergence and vorticity (we note again that the sign of vorticity is changed so that Southern Hemisphere results conform to what is expected in the Northern Hemisphere):
i1520-0426-16-10-1384-e38
Table 2 also contains vorticity and divergence calculations from the wind fields of the Australian Bureau of Meteorology Global Assimilation Prognosis (GASP) NWP model (Bourke et al. 1995). There is little agreement between the two techniques, especially for the divergence, as only one flight shows good agreement for the divergence. The vorticity shows some agreement in six of the nine flights. Furthermore, the average vorticity over the nine flights of −1.8 × 10−5 s−1 is in reasonable agreement with the NWP model average of −2.2 × 10−5 s−1. However, the ratio of the average measured divergence to the average vorticity is about −0.6—considerably smaller in magnitude than that predicted from (31). We anticipate the better agreement for the vorticity as it is also a more robust measurement for calculations based on the global NWP. Equation (39) typically leads to a small difference between two relatively large numbers, while the vorticity (38) is typically a summation, which results in a smaller error.

The NWP-based estimates suffer from additional problems. First, they are particularly sensitive to the initial position and time. Changing the height of the calculations from 500 to 1000 m leads to a change of as much as 3 × 10−6 s−1 in the divergence in some flights. European Centre for Medium-Range Weather Forecasts (ECMWF)-based divergences also varied strongly with altitude for the ACE-1 Lagrangians, as shown in W99a. These divergences may also vary in time as noted by Bretherton et al. (1995) for ECMWF-based calculations for the ASTEX Lagrangians.

Second, the difference in scales between the in situ measurements and the NWP measurements may also be a problem. At best, measurements based on global NWP datasets are smoothed to at least the resolution of the model, typically on the order of 1°. In this region of the globe, however, observations are sparser. The global analyses usually apply only to synoptic-scale variations, which means that smaller-scale structures, such as fronts, are underresolved. Flights 13, 18, and 19 were conducted in the neighborhood of fronts and have poor agreement in vorticity calculations in Table 2. If the divergence or vorticity has a mesoscale variability, we would not be able to see it in the NWP-based calculations. This explains why the in situ calculations of divergence can be an order of magnitude greater than the synoptic-scale values. We also note that there is little advantage in using smaller-scale, nested models to derive the divergence and vorticity since these models still depend on the global analysis for boundary conditions and initialization in this region of the globe.

We can average over some of the temporal and spatial variability by considering the divergence and vorticity observed on numerical trajectories. Here we used the change in altitude of a trajectory to obtain the divergence and the rotation of a cluster of trajectories to obtain the local vorticity. Following S, we consider numerical trajectories calculated from three independent global NWP models: the Australian Bureau of Meteorology GASP model, the National Oceanic and Atmospheric Administration (NOAA) aviation model and the NOAA medium-range forecast model. The trajectories were produced by the HYSPLIT code (Draxler 1991); full details of these calculations are found in S99. For this paper, numerical trajectories were initialized at the center of the first circle for each flight at an elevation of 500 m and allowed to run for 6 h. Trajectories were based on the full three-dimensional winds.

Table 3 shows that the vorticity is reasonably consistent among the different models but the divergence is not. S99 found that the numerical horizontal motions of the trajectories were also consistent among the models. Errors typically were 5%–15% of the total trajectory length. The vertical motion of their trajectories, however, varied among the models so that there was little confidence in the divergence derived from global NWP models.

Finally we illustrate the calculation of the entrainment rate We using the divergence from a particular ACE-1 case. Boers et al. (1998) have already derived the divergences for flight 12, which they described as the best example of a classic well-mixed, cloud-topped marine PBL among the ACE-1 flights. They used three different methods to calculate the entrainment rates including Eq. (11). They also discuss the uncertainty in each technique and sources of potential error. First Boers et al. calculated the entrainment rate from the evolution of the thermodynamic budgets of the PBL (Bretherton et al. 1995). Using the liquid water potential temperature, they obtain We = 0.004 m s−1; using the total water mixing ratio, We = 0.006 m s−1. These calculations require estimates of surface fluxes, radiative exchange through the PBL and precipitation, all of which have inherent uncertainties. They also assume simple averages of quantities across the PBL, which sometimes may be inadequate because of the wind shear.

Next Boers et al. calculated We from (2) using ozone as a tracer and obtained We = 0.0033 m s−1. However, there may be large uncertainties associated with both the jump across the inversion and the flux at cloud top (Bretherton et al. 1995). Finally, Boers et al. calculated the entrainment directly from Eq. (2). They found virtually no change in zi, which was ∼1300 m over the course of the flight. Using global NWP data they found a divergence of 2.7 × 10−6 s−1; thus We = 0.0035 m s−1. In addition to the uncertainty in the divergence, the change in the inversion height was based on C-130 soundings. This method has a considerable uncertainty due to variations in cloud-top height. Examining the lidar legs for these flights confirms that the height of the inversion was unchanging, although it seemed to be closer to 1100 m. In Boers et al., this last technique is no more accurate than either of the first two. The divergence they use, 2.7 × 10−6 s−1, is roughly three times greater than the GASP values we have in Table 3 because of using different levels for their calculations. Using the circles to calculate the divergence (Table 2) leads to a divergence of 5.1 × 10−6 s−1. This is approximately twice that calculated by Boers et al., so that our entrainment velocity would be twice what they derived.

4. Conclusions

We have discussed three techniques for measuring entrainment at the top of the PBL from an aircraft using the same flight pattern: the budget technique, the flux/δS technique, and the divergence technique. The current limitations in measurement accuracy (or our ignorance in knowing what can be achieved with existing instrumentation) of the lateral wind component mean that the flux/δS approach, which has already been tested and implemented, seems at present to be the most accurate of the three techniques, assuming a suitable tracer can be used. The ideal characteristics are a uniform surface source, a lifetime of 104 < τ < 106 s, and negligibly soluble in cloud. The limited lifetime helps to ensure that there is a measurable flux near the interface and that the jump in concentration across the interface is measurable. However, if entrainment occurs in both directions across an interface (e.g., Russell et al. 1999), this approach requires measurement of the flux on both sides of the interface.

We have shown that the divergence technique has the potential to measure mean vertical velocity, and hence, with concurrent time-resolved measurements of zi, the net entrainment velocity across the PBL top. However, the measurements need to be carried out in a horizontally homogeneous region with a well-mixed PBL over a period of several hours. Another possible limitation of the divergence technique is significant time changes occurring in the flow while flying around the closed path. This possibility argues for a smaller closed path, while the measurement accuracy limitations argue for a larger path; estimating the optimal size requires further analysis.

Although we have assumed a well-mixed PBL for the flux/δS approach, strictly speaking this is not essential for the technique to work. A basic requirement is a flux measurement at the top of a well-mixed layer, with an abrupt transition to a nonturbulent fluid above. Measuring turbulent flux from an aircraft very close to this interface is difficult since the aircraft needs to stay within the well-mixed layer to avoid sampling errors due to large changes in both scalar quantities and velocity that can occur across the interface. Typically variations in height of tens to a hundred meters or more can occur at the top of well-mixed layers. This is why the flux at the interface is best estimated by extrapolating from flux measurements at lower levels. Thus, if the well-mixed layer is less than a few hundred meters, it may be difficult to estimate the flux at the transition.

The divergence technique does not have that limitation. On the other hand, in a “decoupled PBL”—that is, where turbulent mixing is not continuous throughout the layer—the horizontal wind field may also not be well mixed vertically, which means that the divergence and vorticity are more likely to vary in a nonsimple way with height. Finally, although we have obtained evidence that divergence and vorticity can be measured by this technique, we believe that to demonstrate more rigorously its accuracy and limitations it is necessary to carry out a focused experiment. For example, repeated circles flown at the same level would be useful for estimating the repeatability of the measurement and thus for comparison with theoretical estimates of the measurement error. Flying two aircraft concurrently would help to quantify temporal changes occurring during the period of measurement. This technique strains currently used instrumentation to the limits of its capabilities. It is likely that more accurate air velocity measurements can be attained with technology that is now available, but not yet in operational use, and its application to this technique could make divergence and vorticity a routine measurement from research aircraft in the future.

Acknowledgments

This research is a contribution to the International Global Atmospheric Chemistry (IGAC) Core project of the International Geosphere–Biosphere Programme (IGBP) and is part of the IGAC Aerosol Characterization Experiments (ACE). We thank Bob Gall for supporting the participation of DHL in ACE-1, and Barry Huebert and Alan Bandy for their willingness to listen, ask questions, and take a chance on a newfangled approach to airplane deployment for ACE-1. We are very appreciative of the thorough review provided by Leif Kristensen. We also thank Jielun Sun and Larry Mahrt for their comments. Participation of PBK in this experiment was made possible through financial support from the Cape Grim Baseline Air Pollution Station.

REFERENCES

  • Batchelor, G. K., 1953: Homogeneous Turbulence. Cambridge University Press, 208 pp.

  • Bates, T. S., J. D. Cline, R. H. Gammon, and S. R. Kelly-Hansen, 1987:Regional and seasonal variations in the flux of oceanic dimethyl sulfide to the atmosphere. J. Geophys. Res.,92, 2930–2938.

    • Crossref
    • Export Citation
  • ——, B. J. Huebert, J. L. Gras, B. Griffiths, and P. A. Durkee, 1999:The International Global Atmospheric Chemistry (IGAC) project’s first Aerosol Characterization Experiment (ACE-1)—An overview. J. Geophys. Res., in press.

  • Bendat, J. S., and A. G. Piersol, 1971: Random Data: Analysis and Measurement Procedures. Wiley-Interscience, 407 pp.

  • Boers, R., P. B. Krummel, S. T. Siems, and G. Hess, 1999: Thermodynamic structure and entrainment of stratocumulus over the Southern Ocean, 1998. J. Geophys. Res., in press.

    • Crossref
    • Export Citation
  • Bourke, W., T. Hart, P. Steinle, R. Seaman, G. Embery, M. Naughton, and L. Rikus, 1995: Evolution of the Bureau of Meteorology Global Assimilation and Prediction system. Part 2: Resolution enhancements and case studies. Aust. Meteor. Mag.,44, 19–40.

  • Bretherton, C. S., P. Austin, and S. T. Siems, 1995: Cloudiness and marine boundary layer dynamics in the ASTEX Lagrangian experiments. Part II: Cloudiness, drizzle, surface fluxes, and entrainment. J. Atmos. Sci.,52, 2724–2735.

  • Businger, S., R. Johnson, J. Katzfey, S. T. Siems, and Q. Wang, 1999:Smart tetroons for Lagrangian air-mass tracking during ACE-1. J. Geophys. Res.,104, 11 709–11 722.

    • Crossref
    • Export Citation
  • Ching, J. K. S., 1975: Determining the drag coefficient from vorticity, momentum, and mass budget analyses. J. Atmos. Sci.,32, 1898–1908.

    • Crossref
    • Export Citation
  • Cooper, W. A., and S. Shertz, 1995: An intermittent-sampling system for measuring fluxes from aircraft. Preprints, Ninth Symp. on Meteorological Observations and Instrumentation, Charlotte, NC, Amer. Meteor. Soc., 189–194.

  • Davis, K. J., 1992: Surface fluxes of trace gases derived from convective-layer profiles. Ph.D. dissertation, University of Colorado, 281 pp. [Available as NCAR/CT-139 from NCAR, P. O. Box 3000, Boulder, CO 80307.].

  • ——, D. H. Lenschow, and P. R. Zimmerman, 1994: Biogenic non-methane hydrocarbon emissions estimated from tethered balloon observations. J. Geophys. Res.,99 (D12), 25 587–25 598.

    • Crossref
    • Export Citation
  • ——, ——, S. P. Oncley, C. Kiemle, G. Ehret, A. Giez, and J. Mann, 1997: The role of entrainment in surface–atmosphere interactions over the boreal forest. J. Geophys. Res.,102 (D24), 29 219–29 230.

    • Crossref
    • Export Citation
  • Draxler, R. R., 1991: The accuracy of trajectories during ANATEX calculated using dynamical model analyses versus rawinsonde observations. J. Appl. Meteor.,30, 1446–1467.

    • Crossref
    • Export Citation
  • Hills, A. J., D. H. Lenschow, and J. W. Birks, 1998: Fast dimethyl sulfide measurement by fluorine-induced chemiluminescence. Anal. Chem.,70, 1735–1742.

    • Crossref
    • Export Citation
  • Holton, J. R., 1972: An Introduction to Dynamic Meteorology. Academic Press, 511 pp.

  • Kawa, S. R., and R. Pearson Jr., 1989a: An observational study of stratocumulus entrainment and thermodynamics. J. Atmos. Sci.,46, 2649–2661.

    • Crossref
    • Export Citation
  • ——, and ——, 1989b: Ozone budgets form the Dynamics and Chemistry of Marine Stratocumulus Experiment. J. Geophys. Res.,94, 9809–9817.

    • Crossref
    • Export Citation
  • Keeler, R. J., R. J. Serafin, R. L. Serafin, D. H. Lenschow, J. M. Vaughan, and A. A. Woodfield, 1987: An airborne laser air motion sensing system. Part I: Concept and preliminary experiment. J. Atmos. Oceanic Technol.,4, 113–127.

    • Crossref
    • Export Citation
  • Klein, S. A., and D. L. Hartmann, 1993: The seasonal cycle of low stratiform cloud. J. Climate,6, 1587–1606.

    • Crossref
    • Export Citation
  • Kristensen, L., and D. H. Lenschow, 1987: An airborne laser air motion sensing system. Part II: Design criteria and measurement possibilities. J. Atmos. Oceanic Technol.,4, 128–138.

    • Crossref
    • Export Citation
  • ——, P. Kirkegaard, and D. H. Lenschow, 1983: Squashed atmospheric turbulence. Risø Rep. R-478, Risø. National Laboratory, Roskilde, Denmark, 79 pp. [Available from Risø National Laboratory, DK-4000, Roskilde, Denmark.].

  • Lenschow, D. H., 1990: Factors affecting the structure and stability of boundary-layer clouds. Preprints, Conf. on Cloud Physics, San Francisco, CA, Amer. Meteor. Soc., 37–42.

  • ——, 1995: Micrometeorological techniques for measuring biosphere–atmosphere trace gas exchange. Biogenic Trace Gases: Measuring Emissions from Soil and Water, P. A. Matson and R. C. Harriss, Ed., Blackwell Science, 126–163.

  • ——, 1996: A proposal for measuring entrainment into the cloud-capped boundary layer. Proc. ETL/CSU Cloud-Related Process Modeling and Measurement Workshop, Boulder, CO, NOAA/ETL, 29–55.

  • ——, and L. Kristensen, 1985: Uncorrelated noise in turbulence measurements. J. Atmos. Oceanic Technol.,2, 68–81.

    • Crossref
    • Export Citation
  • ——, and B. B. Stankov, 1986: Length scales in the convective atmospheric boundary layer. J. Atmos. Sci.,43, 1198–1209.

    • Crossref
    • Export Citation
  • ——, and L. Kristensen, 1988: Applications of dual aircraft formation flights. J. Atmos. Oceanic Technol.,5, 715–726.

    • Crossref
    • Export Citation
  • ——, J. C. Wyngaard, and W. T. Pennell, 1980: Mean-field and second-moments budgets in a baroclinic, convective boundary layer. J. Atmos. Sci.,37, 1313–1326.

    • Crossref
    • Export Citation
  • ——, R. Pearson Jr., and B. B. Stankov, 1981: Estimating the ozone budget in the boundary layer by use of aircraft measurements of ozone eddy flux and mean concentration. J. Geophys. Res.,86, 7291–7297.

    • Crossref
    • Export Citation
  • ——, J. Mann, and L. Kristensen, 1994: How long is long enough when measuring fluxes and other turbulence statistics? J. Atmos. Oceanic Technol.,11, 661–673.

    • Crossref
    • Export Citation
  • ——, I. R. Paluch, A. R. Bandy, D. C. Thornton, D. R. Blake, and I. Simpson, 1999: Use of a mixed-layer model to estimate dimethyl sulfide flux and application to other trace gas species. J. Geophys. Res., in press.

  • Lilly, D. K., 1968: Models of cloud-topped mixed layers under a strong inversion. Quart. J. Roy. Meteor. Soc.,94, 292–309.

    • Crossref
    • Export Citation
  • Mann, J., and D. H. Lenschow, 1994: Errors in airborne flux measurement. J. Geophys. Res.,99 (D7), 14 519–14 526.

    • Crossref
    • Export Citation
  • Moeng, C.-H., and J. C. Wyngaard, 1984: Statistics of conservative scalars in the convective boundary layer. J. Atmos. Sci.,41, 3161–3169.

    • Crossref
    • Export Citation
  • ——, and ——, 1989: Evaluation of turbulent transport and dissipation closures in second-order modeling. J. Atmos. Sci.,46, 2311–2330.

    • Crossref
    • Export Citation
  • ——, and Coauthors, 1996: Simulation of a stratocumulus-topped PBL: Intercomparison among different numerical codes. Bull. Amer. Meteor. Soc.,77, 261–278.

    • Crossref
    • Export Citation
  • Panofsky, H. A., and J. A. Dutton, 1984: Atmospheric Turbulence. Wiley and Sons, 397 pp.

  • Piper, M., J. C. Wyngaard, W. H. Snyder, and R. E. Lawson Jr., 1995:Convection tank experiments on top–down, bottom–up diffusion. Preprints, 11th Symp. on Boundary Layers and Turbulence, Charlotte, NC, Amer. Meteor. Soc., 59–62.

  • Russell, L. M., D. H. Lenschow, K. K. Laursen, P. B. Krummel, S. T. Siems, A. Bandy, D. Thornton, and T. S. Bates, 1998: Bidirectional mixing in an ACE-1 marine PBL overlain by a second turbulent layer. J. Geophys. Res.,103, 16 411–16 432.

    • Crossref
    • Export Citation
  • Schubert, W. H., F. S. Wakefield, E. J. Steiner, and S. K. Cox, 1979:Marine stratocumulus convection. Part 1: Governing equations and horizontally homogeneous solutions. J. Atmos. Sci.,36, 1286–1307.

  • Siems, S. T., G. D. Hess, K. Suhre, S. Businger, and R. R. Draxler, 1999: The impact of wind shear on observed and simulated trajectories during the ACE-1 Lagrangian Experiment. J. Geophys. Res., in press.

  • Wang, Q., and Coauthors, 1999a: Characteristics of the marine boundary layer observed during Lagrangian measurement. Part 1: General conditions and mean characteristics. J. Geophys. Res., in press.

    • Crossref
    • Export Citation
  • ——, and Coauthors, 1999b: Characteristics of the marine boundary layer observed during Lagrangian measurements. Part 2: Turbulence structure. J. Geophys. Res., in press.

  • Wyngaard, J. C., and R. A. Brost, 1984: Top–down and bottom–up diffusion of a scalar in the convective boundary layer. J. Atmos. Sci.,41, 102–112.

    • Crossref
    • Export Citation

Fig. 1.
Fig. 1.

Normalized error variance as a function of normalized height z∗ = z/zi for measurement of a scalar flux in the convective marine boundary layer. The numbers on the curves denote different values of α = (ws)0/(ws)zi.

Citation: Journal of Atmospheric and Oceanic Technology 16, 10; 10.1175/1520-0426(1999)016<1384:MEDAVO>2.0.CO;2

Fig. 2.
Fig. 2.

Autocorrelation function for a horizontal wind component in the vertical direction plotted in terms of the normalized separation distance ζ/zi.

Citation: Journal of Atmospheric and Oceanic Technology 16, 10; 10.1175/1520-0426(1999)016<1384:MEDAVO>2.0.CO;2

Fig. 3.
Fig. 3.

(a) Schematic of a horizontal velocity field in an earth-based coordinate system, which increases linearly in the x direction showing the trajectory of an airplane in a constant turn rate. (b) Schematic of the horizontal velocity field (a) after a Galilean transformation has removed the mean horizontal air velocity seen by the aircraft. The flight track now is closed and nearly circular with radius R. The velocity component normal to the aircraft flight track υn is also shown, with an error ε in its direction.

Citation: Journal of Atmospheric and Oceanic Technology 16, 10; 10.1175/1520-0426(1999)016<1384:MEDAVO>2.0.CO;2

Fig. 4.
Fig. 4.

(a) Pressure altitude of the NCAR C-130 vs time for flight 18 (1 Dec 1995), with the levels of the circles shown in the lower panel numbered and (b) horizontal track of the aircraft.

Citation: Journal of Atmospheric and Oceanic Technology 16, 10; 10.1175/1520-0426(1999)016<1384:MEDAVO>2.0.CO;2

Fig. 5.
Fig. 5.

(a) Divergences calculated from the 10 30-min circles of flight 18 plotted as a function of pressure. (b) Vorticity calculated from the 10 30-min circles of flight 18 plotted as a function of pressure. Here and in subsequent figures, the vorticity sign has been reversed to conform to what would be expected in the Northern Hemisphere.

Citation: Journal of Atmospheric and Oceanic Technology 16, 10; 10.1175/1520-0426(1999)016<1384:MEDAVO>2.0.CO;2

Fig. 6.
Fig. 6.

An example of a circular flight path from flight 18. The trajectory is corrected for advection by the mean wind measured by the aircraft. The square encloses the starting and ending points (separated by 175 m); the ending point is obtained by determining the minimum distance from the starting point. The crosses on either side of the square are ±20 s from the end point, and the cross farther away on the left side is at 60 s from the end point.

Citation: Journal of Atmospheric and Oceanic Technology 16, 10; 10.1175/1520-0426(1999)016<1384:MEDAVO>2.0.CO;2

Fig. 7.
Fig. 7.

Smoothly varying lines are running calculations of the integrals of the wind components normal to the aircraft track (divergence) and along the aircraft track (vorticity) normalized by the fractional area of the circle [πR2 × θ/(2π)], where θ is the angle subtended by the aircraft trajectory. The wind components themselves are the lines with considerable small-scale variability. This calculation is for the circle in Fig. 6.

Citation: Journal of Atmospheric and Oceanic Technology 16, 10; 10.1175/1520-0426(1999)016<1384:MEDAVO>2.0.CO;2

Fig. 8.
Fig. 8.

The final few minutes of the running divergence and vorticity calculation shown in Fig. 7, indicating the sensitivity of the calculation to the accuracy in estimating the closure of the circle.

Citation: Journal of Atmospheric and Oceanic Technology 16, 10; 10.1175/1520-0426(1999)016<1384:MEDAVO>2.0.CO;2

Fig. 9.
Fig. 9.

The autocorrelation function and integral of the autocorrelation function for the first circle of flight 18. The maximum of the integrated autocorrelation function is used as an estimate of the integral scale.

Citation: Journal of Atmospheric and Oceanic Technology 16, 10; 10.1175/1520-0426(1999)016<1384:MEDAVO>2.0.CO;2

Table 1.

Estimates of the changes in divergence and vorticity as a function of time offset or closure distance. For a time offset of 0 s, the end point or “closure point” is the minimum distance between the starting point and the flight track near the end of the circle.

Table 1.
Table 2.

Summary of mean divergences and vorticities from all usable ACE-1 flights with at least four PBL circles. Positive vorticityindicates cyclonic or clockwise rotation in the Southern Hemisphere.

Table 2.
Table 3.

Comparison among divergences and vorticities obtained from different analyses models.

Table 3.

* The National Center for Atmospheric Research is sponsored by the National Science Foundation.

Save
  • Batchelor, G. K., 1953: Homogeneous Turbulence. Cambridge University Press, 208 pp.

  • Bates, T. S., J. D. Cline, R. H. Gammon, and S. R. Kelly-Hansen, 1987:Regional and seasonal variations in the flux of oceanic dimethyl sulfide to the atmosphere. J. Geophys. Res.,92, 2930–2938.

    • Crossref
    • Export Citation
  • ——, B. J. Huebert, J. L. Gras, B. Griffiths, and P. A. Durkee, 1999:The International Global Atmospheric Chemistry (IGAC) project’s first Aerosol Characterization Experiment (ACE-1)—An overview. J. Geophys. Res., in press.

  • Bendat, J. S., and A. G. Piersol, 1971: Random Data: Analysis and Measurement Procedures. Wiley-Interscience, 407 pp.

  • Boers, R., P. B. Krummel, S. T. Siems, and G. Hess, 1999: Thermodynamic structure and entrainment of stratocumulus over the Southern Ocean, 1998. J. Geophys. Res., in press.

    • Crossref
    • Export Citation
  • Bourke, W., T. Hart, P. Steinle, R. Seaman, G. Embery, M. Naughton, and L. Rikus, 1995: Evolution of the Bureau of Meteorology Global Assimilation and Prediction system. Part 2: Resolution enhancements and case studies. Aust. Meteor. Mag.,44, 19–40.

  • Bretherton, C. S., P. Austin, and S. T. Siems, 1995: Cloudiness and marine boundary layer dynamics in the ASTEX Lagrangian experiments. Part II: Cloudiness, drizzle, surface fluxes, and entrainment. J. Atmos. Sci.,52, 2724–2735.

  • Businger, S., R. Johnson, J. Katzfey, S. T. Siems, and Q. Wang, 1999:Smart tetroons for Lagrangian air-mass tracking during ACE-1. J. Geophys. Res.,104, 11 709–11 722.

    • Crossref
    • Export Citation
  • Ching, J. K. S., 1975: Determining the drag coefficient from vorticity, momentum, and mass budget analyses. J. Atmos. Sci.,32, 1898–1908.

    • Crossref
    • Export Citation
  • Cooper, W. A., and S. Shertz, 1995: An intermittent-sampling system for measuring fluxes from aircraft. Preprints, Ninth Symp. on Meteorological Observations and Instrumentation, Charlotte, NC, Amer. Meteor. Soc., 189–194.

  • Davis, K. J., 1992: Surface fluxes of trace gases derived from convective-layer profiles. Ph.D. dissertation, University of Colorado, 281 pp. [Available as NCAR/CT-139 from NCAR, P. O. Box 3000, Boulder, CO 80307.].

  • ——, D. H. Lenschow, and P. R. Zimmerman, 1994: Biogenic non-methane hydrocarbon emissions estimated from tethered balloon observations. J. Geophys. Res.,99 (D12), 25 587–25 598.

    • Crossref
    • Export Citation
  • ——, ——, S. P. Oncley, C. Kiemle, G. Ehret, A. Giez, and J. Mann, 1997: The role of entrainment in surface–atmosphere interactions over the boreal forest. J. Geophys. Res.,102 (D24), 29 219–29 230.

    • Crossref
    • Export Citation
  • Draxler, R. R., 1991: The accuracy of trajectories during ANATEX calculated using dynamical model analyses versus rawinsonde observations. J. Appl. Meteor.,30, 1446–1467.

    • Crossref
    • Export Citation
  • Hills, A. J., D. H. Lenschow, and J. W. Birks, 1998: Fast dimethyl sulfide measurement by fluorine-induced chemiluminescence. Anal. Chem.,70, 1735–1742.

    • Crossref
    • Export Citation
  • Holton, J. R., 1972: An Introduction to Dynamic Meteorology. Academic Press, 511 pp.

  • Kawa, S. R., and R. Pearson Jr., 1989a: An observational study of stratocumulus entrainment and thermodynamics. J. Atmos. Sci.,46, 2649–2661.

    • Crossref
    • Export Citation
  • ——, and ——, 1989b: Ozone budgets form the Dynamics and Chemistry of Marine Stratocumulus Experiment. J. Geophys. Res.,94, 9809–9817.

    • Crossref
    • Export Citation
  • Keeler, R. J., R. J. Serafin, R. L. Serafin, D. H. Lenschow, J. M. Vaughan, and A. A. Woodfield, 1987: An airborne laser air motion sensing system. Part I: Concept and preliminary experiment. J. Atmos. Oceanic Technol.,4, 113–127.

    • Crossref
    • Export Citation
  • Klein, S. A., and D. L. Hartmann, 1993: The seasonal cycle of low stratiform cloud. J. Climate,6, 1587–1606.

    • Crossref
    • Export Citation
  • Kristensen, L., and D. H. Lenschow, 1987: An airborne laser air motion sensing system. Part II: Design criteria and measurement possibilities. J. Atmos. Oceanic Technol.,4, 128–138.

    • Crossref
    • Export Citation
  • ——, P. Kirkegaard, and D. H. Lenschow, 1983: Squashed atmospheric turbulence. Risø Rep. R-478, Risø. National Laboratory, Roskilde, Denmark, 79 pp. [Available from Risø National Laboratory, DK-4000, Roskilde, Denmark.].

  • Lenschow, D. H., 1990: Factors affecting the structure and stability of boundary-layer clouds. Preprints, Conf. on Cloud Physics, San Francisco, CA, Amer. Meteor. Soc., 37–42.

  • ——, 1995: Micrometeorological techniques for measuring biosphere–atmosphere trace gas exchange. Biogenic Trace Gases: Measuring Emissions from Soil and Water, P. A. Matson and R. C. Harriss, Ed., Blackwell Science, 126–163.

  • ——, 1996: A proposal for measuring entrainment into the cloud-capped boundary layer. Proc. ETL/CSU Cloud-Related Process Modeling and Measurement Workshop, Boulder, CO, NOAA/ETL, 29–55.

  • ——, and L. Kristensen, 1985: Uncorrelated noise in turbulence measurements. J. Atmos. Oceanic Technol.,2, 68–81.

    • Crossref
    • Export Citation
  • ——, and B. B. Stankov, 1986: Length scales in the convective atmospheric boundary layer. J. Atmos. Sci.,43, 1198–1209.

    • Crossref
    • Export Citation
  • ——, and L. Kristensen, 1988: Applications of dual aircraft formation flights. J. Atmos. Oceanic Technol.,5, 715–726.

    • Crossref
    • Export Citation
  • ——, J. C. Wyngaard, and W. T. Pennell, 1980: Mean-field and second-moments budgets in a baroclinic, convective boundary layer. J. Atmos. Sci.,37, 1313–1326.

    • Crossref
    • Export Citation
  • ——, R. Pearson Jr., and B. B. Stankov, 1981: Estimating the ozone budget in the boundary layer by use of aircraft measurements of ozone eddy flux and mean concentration. J. Geophys. Res.,86, 7291–7297.

    • Crossref
    • Export Citation
  • ——, J. Mann, and L. Kristensen, 1994: How long is long enough when measuring fluxes and other turbulence statistics? J. Atmos. Oceanic Technol.,11, 661–673.

    • Crossref
    • Export Citation
  • ——, I. R. Paluch, A. R. Bandy, D. C. Thornton, D. R. Blake, and I. Simpson, 1999: Use of a mixed-layer model to estimate dimethyl sulfide flux and application to other trace gas species. J. Geophys. Res., in press.

  • Lilly, D. K., 1968: Models of cloud-topped mixed layers under a strong inversion. Quart. J. Roy. Meteor. Soc.,94, 292–309.

    • Crossref
    • Export Citation
  • Mann, J., and D. H. Lenschow, 1994: Errors in airborne flux measurement. J. Geophys. Res.,99 (D7), 14 519–14 526.

    • Crossref
    • Export Citation
  • Moeng, C.-H., and J. C. Wyngaard, 1984: Statistics of conservative scalars in the convective boundary layer. J. Atmos. Sci.,41, 3161–3169.

    • Crossref
    • Export Citation
  • ——, and ——, 1989: Evaluation of turbulent transport and dissipation closures in second-order modeling. J. Atmos. Sci.,46, 2311–2330.

    • Crossref
    • Export Citation
  • ——, and Coauthors, 1996: Simulation of a stratocumulus-topped PBL: Intercomparison among different numerical codes. Bull. Amer. Meteor. Soc.,77, 261–278.

    • Crossref
    • Export Citation
  • Panofsky, H. A., and J. A. Dutton, 1984: Atmospheric Turbulence. Wiley and Sons, 397 pp.

  • Piper, M., J. C. Wyngaard, W. H. Snyder, and R. E. Lawson Jr., 1995:Convection tank experiments on top–down, bottom–up diffusion. Preprints, 11th Symp. on Boundary Layers and Turbulence, Charlotte, NC, Amer. Meteor. Soc., 59–62.

  • Russell, L. M., D. H. Lenschow, K. K. Laursen, P. B. Krummel, S. T. Siems, A. Bandy, D. Thornton, and T. S. Bates, 1998: Bidirectional mixing in an ACE-1 marine PBL overlain by a second turbulent layer. J. Geophys. Res.,103, 16 411–16 432.

    • Crossref
    • Export Citation
  • Schubert, W. H., F. S. Wakefield, E. J. Steiner, and S. K. Cox, 1979:Marine stratocumulus convection. Part 1: Governing equations and horizontally homogeneous solutions. J. Atmos. Sci.,36, 1286–1307.

  • Siems, S. T., G. D. Hess, K. Suhre, S. Businger, and R. R. Draxler, 1999: The impact of wind shear on observed and simulated trajectories during the ACE-1 Lagrangian Experiment. J. Geophys. Res., in press.

  • Wang, Q., and Coauthors, 1999a: Characteristics of the marine boundary layer observed during Lagrangian measurement. Part 1: General conditions and mean characteristics. J. Geophys. Res., in press.

    • Crossref
    • Export Citation
  • ——, and Coauthors, 1999b: Characteristics of the marine boundary layer observed during Lagrangian measurements. Part 2: Turbulence structure. J. Geophys. Res., in press.

  • Wyngaard, J. C., and R. A. Brost, 1984: Top–down and bottom–up diffusion of a scalar in the convective boundary layer. J. Atmos. Sci.,41, 102–112.

    • Crossref
    • Export Citation
  • Fig. 1.

    Normalized error variance as a function of normalized height z∗ = z/zi for measurement of a scalar flux in the convective marine boundary layer. The numbers on the curves denote different values of α = (ws)0/(ws)zi.

  • Fig. 2.

    Autocorrelation function for a horizontal wind component in the vertical direction plotted in terms of the normalized separation distance ζ/zi.

  • Fig. 3.

    (a) Schematic of a horizontal velocity field in an earth-based coordinate system, which increases linearly in the x direction showing the trajectory of an airplane in a constant turn rate. (b) Schematic of the horizontal velocity field (a) after a Galilean transformation has removed the mean horizontal air velocity seen by the aircraft. The flight track now is closed and nearly circular with radius R. The velocity component normal to the aircraft flight track υn is also shown, with an error ε in its direction.

  • Fig. 4.

    (a) Pressure altitude of the NCAR C-130 vs time for flight 18 (1 Dec 1995), with the levels of the circles shown in the lower panel numbered and (b) horizontal track of the aircraft.

  • Fig. 5.

    (a) Divergences calculated from the 10 30-min circles of flight 18 plotted as a function of pressure. (b) Vorticity calculated from the 10 30-min circles of flight 18 plotted as a function of pressure. Here and in subsequent figures, the vorticity sign has been reversed to conform to what would be expected in the Northern Hemisphere.

  • Fig. 6.

    An example of a circular flight path from flight 18. The trajectory is corrected for advection by the mean wind measured by the aircraft. The square encloses the starting and ending points (separated by 175 m); the ending point is obtained by determining the minimum distance from the starting point. The crosses on either side of the square are ±20 s from the end point, and the cross farther away on the left side is at 60 s from the end point.

  • Fig. 7.

    Smoothly varying lines are running calculations of the integrals of the wind components normal to the aircraft track (divergence) and along the aircraft track (vorticity) normalized by the fractional area of the circle [πR2 × θ/(2π)], where θ is the angle subtended by the aircraft trajectory. The wind components themselves are the lines with considerable small-scale variability. This calculation is for the circle in Fig. 6.

  • Fig. 8.

    The final few minutes of the running divergence and vorticity calculation shown in Fig. 7, indicating the sensitivity of the calculation to the accuracy in estimating the closure of the circle.

  • Fig. 9.

    The autocorrelation function and integral of the autocorrelation function for the first circle of flight 18. The maximum of the integrated autocorrelation function is used as an estimate of the integral scale.

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 1114 607 40
PDF Downloads 299 66 4