Roll and Cell Convection in Wintertime Arctic Cold-Air Outbreaks

Burghard Brümmer Meteorological Institute, University of Hamburg, Hamburg, Germany

Search for other papers by Burghard Brümmer in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

Cold-air outbreaks from the polar ice caps or winterly continents over the open ocean lead to organized convection that typically starts as longitudinal roll patterns and changes to cellular patterns in downstream direction. During the field experiments ARKTIS 1991 and ARKTIS 1993, aircraft missions were conducted in 13 cold-air outbreak events over the Greenland and Barents Seas to determine the characteristic parameters of both the mean (primary) flow and the superimposed organized convection (secondary flow). The measurements are classified into four categories with respect to the convective pattern form: longitudinal rolls with small and wider horizontal wavelengths, transitional forms between rolls and cells, and cells.

Rolls were observed for boundary layer depths h < 1 km with horizontal wavelengths λ < 5 km and aspect ratios λ/h between 2.6 and 6.5. Distinct cellular structures occurred for h > 1.4 km with λ > 8 km and λ/h between 4 and 12. The amplitudes of the secondary flow-scale variations of the temperature θR, moisture mR, and the longitudinal, uR; transversal, υR; and vertical, wR, wind components were on the order of 0.1–0.4 K, 0.03–0.30 g kg−1, 0.6–2.5 m s−1, 0.8–2.5 m s−1, and 0.4–1.8 m s−1, respectively, generally increasing from the roll to the cell region. The same is true for the ratio uR/υR (from about 0.6 to nearly 1) and for the ratio LmR/cpθR (from 0.7 to more than 2), hinting at increasing importance of moisture processes in the cell compared to the roll region.

The importance of the secondary-flow transports of heat and momentum in relation to the total vertical transports increases with height and from rolls to cells. Particularly clear is the vertical profile of the vertical moisture transport mRwR, which exhibits a maximum around cloud base and is on the average related to the surface moisture flux as (mRwR)max = 0.35(mw′)o.

The thermodynamic conditions of the basic flow are characterized by the Rayleigh number Ra, the stability of the capping inversion, and the net condensation rate in the cloud layer. Here Ra is clearly overcritical in the whole cold-air outbreak region; it is around 105 in the roll region and around 2 × 106 in the cell region. The Monin–Obukhov stability parameter does not appear to be suitable measure to distinguish between roll and cell convection. The stability above the boundary layer is about two to three times larger for rolls than for cells. The net condensation in clouds is three times larger in cell than in roll regions and the resulting heating of the boundary layer exceeds that of the surface heat flux in the cell region. The kinematic conditions of the basic flow are characterized by a larger shear of the longitudinal wind component u in the roll than in the cell region. The curvature of the u profile is mostly overcritical in rolls and always subcritical in cells.

The secondary flow-scale kinetic energy Ekin,R is related to Ra. The best least squares fit is given by Ekin,R = 3.7Ra0.4.

Corresponding author address: Dr. Burghard Brümmer, Meteorologisches Institut, Universitat Hamburg, Bundesstraße 55, D-20146 Hamburg, Germany.

Abstract

Cold-air outbreaks from the polar ice caps or winterly continents over the open ocean lead to organized convection that typically starts as longitudinal roll patterns and changes to cellular patterns in downstream direction. During the field experiments ARKTIS 1991 and ARKTIS 1993, aircraft missions were conducted in 13 cold-air outbreak events over the Greenland and Barents Seas to determine the characteristic parameters of both the mean (primary) flow and the superimposed organized convection (secondary flow). The measurements are classified into four categories with respect to the convective pattern form: longitudinal rolls with small and wider horizontal wavelengths, transitional forms between rolls and cells, and cells.

Rolls were observed for boundary layer depths h < 1 km with horizontal wavelengths λ < 5 km and aspect ratios λ/h between 2.6 and 6.5. Distinct cellular structures occurred for h > 1.4 km with λ > 8 km and λ/h between 4 and 12. The amplitudes of the secondary flow-scale variations of the temperature θR, moisture mR, and the longitudinal, uR; transversal, υR; and vertical, wR, wind components were on the order of 0.1–0.4 K, 0.03–0.30 g kg−1, 0.6–2.5 m s−1, 0.8–2.5 m s−1, and 0.4–1.8 m s−1, respectively, generally increasing from the roll to the cell region. The same is true for the ratio uR/υR (from about 0.6 to nearly 1) and for the ratio LmR/cpθR (from 0.7 to more than 2), hinting at increasing importance of moisture processes in the cell compared to the roll region.

The importance of the secondary-flow transports of heat and momentum in relation to the total vertical transports increases with height and from rolls to cells. Particularly clear is the vertical profile of the vertical moisture transport mRwR, which exhibits a maximum around cloud base and is on the average related to the surface moisture flux as (mRwR)max = 0.35(mw′)o.

The thermodynamic conditions of the basic flow are characterized by the Rayleigh number Ra, the stability of the capping inversion, and the net condensation rate in the cloud layer. Here Ra is clearly overcritical in the whole cold-air outbreak region; it is around 105 in the roll region and around 2 × 106 in the cell region. The Monin–Obukhov stability parameter does not appear to be suitable measure to distinguish between roll and cell convection. The stability above the boundary layer is about two to three times larger for rolls than for cells. The net condensation in clouds is three times larger in cell than in roll regions and the resulting heating of the boundary layer exceeds that of the surface heat flux in the cell region. The kinematic conditions of the basic flow are characterized by a larger shear of the longitudinal wind component u in the roll than in the cell region. The curvature of the u profile is mostly overcritical in rolls and always subcritical in cells.

The secondary flow-scale kinetic energy Ekin,R is related to Ra. The best least squares fit is given by Ekin,R = 3.7Ra0.4.

Corresponding author address: Dr. Burghard Brümmer, Meteorologisches Institut, Universitat Hamburg, Bundesstraße 55, D-20146 Hamburg, Germany.

1. Introduction

Atmospheric convection organized in longitudinal or cellular patterns is a frequent phenomenon in cold-air outbreaks from the polar ice caps or winterly cold continents over the open water. It is also, but not as frequently, observed in cold-air masses moving over heated land surfaces during daytime in summer. The motion patterns of the inherent secondary flow superimposed to the basic (primary) flow are made visible by cloud development in an often impressive manner on satellite images (Fig. 1). Many experimental, theoretical, and numerical studies on organized convection patterns have been made in the last 30 years. For references see, for example, the review article of Atkinson and Zhang (1996), the review articles of Brown (1980) and Etling and Brown (1993) on longitudinal convection patterns, and the review papers of Agee (1984 1987) on cellular convection patterns.

The driving mechanisms underlying these organized convection patterns (secondary flow) are thermal or dynamical instabilities of the basic flow or a combination of both instabilities. Thermal forcing may be due to heating from the surface, to condensational heating, or to radiational and evaporational cooling at cloud top. Dynamical forcing is due to wind shear in connection with an inflection point in the vertical wind profile. In the above-mentioned situations of cold-air flows over warmer surfaces the thermal instability is the primary forcing and the dynamical instability plays only a minor role. However, vertical shear (e.g., Asai 1970) and curvature (Kuettner 1971) of the basic flow play an important role for the organization of the convective pattern (rolls or cells).

Linear theories (e.g., Lilly 1966; Etling 1971; Kuettner 1971; Rosmond 1973; Helfand and Kalnay 1983) of thermal and dynamical instabilities are able to give information about the geometrical characteristics of the evolving secondary flows, such as wavelength λ, vertical depth h, and aspect ratio λ/h. They also give information about critical values of the relevant numbers that characterize the stability conditions of the basic flow, such as Rayleigh number Ra or Reynolds number Re. Linear theories, however, are not able to determine the absolute amplitudes of the secondary flow variables, such as temperature T, specific humidity q, or the wind components u, υ, and w.

Special observations (e.g., Hein and Brown 1988; Chou and Zimmerman 1989; Wayland and Raman 1989;Chou and Ferguson 1991), nonlinear analytical studies (e.g., Brown 1970; 1972; Brown and Liu 1982), and numerical model studies (e.g., Mason 1989; Raasch 1990; Chlond 1992) indicate that organized convective flows can contribute substantially to the vertical fluxes of heat, moisture, and momentum in the atmospheric boundary layer. In weather and climate models these subgrid-scale fluxes are not accounted for especially (see, e.g., Brown and Foster 1994). Their effects have to be overtaken somehow in these models either by the turbulence or by the convection routines. The latter does not distinguish between organized or randomly distributed convection. It is not known so far whether the transports by organized or randomly distributed convection under otherwise nearly similar conditions are significantly different. Before taking organized convection into account in the convection parameterization schemes, more systematic knowledge about the secondary flow characteristics, that is, geometrical as well as amplitude characteristics, has to be gathered. To make a contribution to this objective is the purpose of this paper. Based on in situ measurements in 13 cases of organized convection in cold-air outbreaks from the Arctic sea ice over the open water, the characteristics of the secondary flow under longitudinal and cellular organization structures have been analyzed and will be presented in this paper. Beyond that, an attempt will be untertaken to relate the secondary flow quantities to the relevant parameters of the basic flow.

2. The data

The data used in this study were measured during the field experiments ARKTIS 1991 and ARKTIS 1993, which took place in February and March 1991 over the Norwegian Sea to the north of the Lofoten Islands and in March 1993 over the West Spitsbergen Current (Fig. 2). A description of the experiments can be found in the field phase reports of Brümmer (1992, 1993).

During both experiments aircraft missions have been conducted to measure the boundary layer modification and the convective pattern transition over the open water during cold-air outbreaks from the Arctic sea ice. In this paper, results from 13 aircraft missions will be presented; nine cases were observed during ARKTIS 1993 and four cases during ARKTIS 1991. A typical flight pattern of an aircraft mission during the ARKTIS 1993 experiment is shown in Fig. 3. Two German research aircraft participated in a flight mission: a Falcon-20 and a Dornier-128. In this paper, we use only the Falcon measurements and therefrom only those parts of the flight patterns that were located over water. Horizontal flight legs perpendicular to the mean wind were flown at several levels in the boundary layer. Data from these legs are used to analyze the geometrical properties of the organized convection patterns as well as the variances and vertical fluxes. Vertical profile flights normally took place at both sides of the vertical plane in which the horizontal crosswind flight legs were embedded. From these profiles those data were extracted that characterize the basic flow properties.

The research aircraft Falcon was equipped with a gust probe system (five-hole flow-angle sensor) to measure the three wind components, with a fast temperature sensor (Rosemount) as well as a slow (Väisälä humicap) and a fast (Lyman α) humidity sensor. The sampling rate was 100 times per second. Furthermore, the Falcon was equipped with an infrared radiometer (Barnes PRT-6) for surface temperature measurement and with instruments to measure the shortwave and longwave radiation fluxes (Eppley: PSP and PIR) from above and below. Data of the radiation sensors were sampled with a rate of 10 times per second.

3. Methods to analyze the parameters of thebasic and secondary flow

a. Basic flow analysis

We refer to a Cartesian coordinate system in which the x axis is parallel to the mean wind direction DD in the boundary layer and the y axis is perpendicular to it. This coordinate system has the advantage that the x axis is nearly aligned with the rotation axis in case of roll convection.

In order to determine the underlying thermal and dynamical mechanisms forcing organized convection we assume that the basic flow is sufficiently characterized by the following parameters: the height h of the boundary layer; the depth Δhc of the cloud layer; the vertical averages of potential temperature θ̃; specific humidity ; longitudinal wind component ũ and lateral wind component υ̃ over the depth h of the boundary layer; the air–sea differences Δθas, Δqas, Δuas, and Δυas and the vertical differences Δθbl, Δqbl, Δubl, and Δυbl across the boundary layer; the vertical differences Δθct, Δqct, Δuct, and Δυct over a 300-m-deep layer above h; the curvature ∂2u/∂z2 and ∂2υ/∂z2 within the boundary layer; and an eventually existing inflection point in the u- or the υ-wind profile with corresponding values for the height huip and hυip.

These parameters were taken from the vertical profiles measured at both sides of the crosswind vertical planes (see Fig. 3). As an example, Fig. 4 shows a profile measured on 24 March 1993 at 1209 UTC at a distance Δx of 160 km from the ice edge and illustrates how the characteristic parameters were specified in this case.

Cloud base hcb and cloud top hct were observed by eye by the aircraft scientist (the author was the aircraft scientist on board the research aircraft Falcon during all flights). Here, we assume that h and hct are identical. The air–sea differences were taken between the sea surface and the lowest permitted flight level at 90 m. The air immediately at the sea surface is assumed to be saturated with respect to the sea surface temperature Ts, that is, qs = qsat(Ts), and to have zero velocity, that is, us = υs = 0.

b. Secondary flow analysis

Parameters representing the characteristics of the secondary flow are analyzed using the data from the crosswind horizontal flight legs (see Fig. 3). These parameters are the horizontal wavelength λ of the rolls (or diameter in case of cells), the variances a2R (a = θ, q, u, υ, w) in the wavelength range Δλ representing the secondary flow, and the vertical fluxes wRaR (a = θ, q, u, υ), where w is the vertical wind component and the index R refers to the secondary flow scale.

As an example, Fig. 5 shows the time series of the downwelling longwave radiation L↓, the temperature T, the water vapor mixing ratio m, and the three wind components u, υ, and w measured during a crosswind flight leg at 92 m on 24 March 1993. The leg was situated directly to the west of the position of the profile displayed in Fig. 4. The aircraft flew from east to west in the direction of the negative y axis at a speed of 100 m s−1 so that the time series covers a total distance of about 42 km. The maxima of L↓ mark the overlying cloud streets, which have wavelengths between 3.2 and 5.0 km. The corresponding secondary flow is clearly reflected in the local wind and thermodynamic field, particularly in the lateral wind component υ and in the mixing ratio m. Sections with confluent lateral flow, ∂υ/∂y < 0, coincide with higher values of m and T and with clouds above. The longitudinal wind component u exhibits also a regular pattern and is lower underneath and higher between the cloud streets. The secondary flow vertical motion w is not that clearly visible because its amplitude is relatively small at this low level, and turbulent as well as cloud-scale updrafts and downdrafts are superimposed. The maximum amplitude of wR occurs normally in the middle of the boundary layer, as will be shown below.

Since the secondary flow pattern is clearly reflected in the cloud field, measurements of the longwave radiation are used to determine the wavelength range Δλ of the secondary flow. Figure 6 shows four variance spectra of longwave radiation calculated from crosswind flight legs flown at 92, 272, 468, and 694 m in the same vertical plane. The spectrum for 92-m height belongs to the time series shown in Fig. 5. Since the cloud pattern is independent of the measuring level in the vertical plane and does not change significantly during the typical duration of the measurements in the same vertical plane (in this case from 1135 to 1208 UTC for four legs), the wavelength range Δλ is fixed by taking all spectra in the same vertical plane into consideration. The limiting frequencies f1 and f2 are chosen as to contain the relevant variance maxima of downwelling (L↓) and upwelling (L↑) longwave radiation. Of course, the choice of f1 and f2 is somewhat subjective, but we have tested that eligible shifts of f1 and f2 do not change the results for the secondary flow-scale variances a2R and fluxes wRaR essentially.

An example of frequency-weighted variance spectra, Saaf, and secondary flow-scale variances, a2R, (a = θ, m, u, υ, w), is presented in Fig. 7. It is calculated by FFT analysis methods from the time series in Fig. 5 and the secondary flow wavelength range Δλ is taken from the limiting frequencies f1 and f2 in Fig. 6. The secondary flow signal appears clearly in all spectra, except in w for the above-mentioned reasons.

Variance and covariance spectra have been calculated for all crosswind flight legs and then integrated both over the entire and the secondary flow-scale frequency range to obtain the total (all frequencies) and secondary flow-scale variances and covariances. If the variations of the secondary flow in the crosswind direction are assumed to be sinusoidal, a = A0 sin[(2π/λ)y + ϕ] (what is close to reality), the following relation between the secondary flow-scale variance a2R and the amplitude A0 of the sinusoidal variation holds:
i1520-0469-56-15-2613-e1
From (1) the average amplitude of the secondary flow-scale variations is estimated.

4. Characteristics of the secondary flow

a. Geometrical dimensions of the secondaryflow patterns

Wavelength λ and vertical depth h of the secondary flow patterns were measured, as mentioned above, in 13 aircraft missions during the field experiments ARKTIS 1991 and ARKTIS 1993. During seven flight missions measurements within a vertical crosswind plane were made at two different downwind distances Δx from the ice edge. All 20 cases are listed in Table 1, showing day, Δx, the geometrical dimensions h, λ, and the aspect ratio λ/h of the secondary flow patterns, as well as the geometrical dimensions of the cloud field, that is, cloud base hcb, cloud layer depth Δhc, and cloud coverage N. Based on the eye observations of the aircraft scientist, on a cockpit video camera, and on satellite images, the cases are classified into four categories with respect to the convective pattern. It is distinguished between rolls (R), transitional plan forms between rolls and cells (R/C), and cells (C). The “roll” class is subdivided further with respect to ice edge distance Δx into two classes named “small” rolls (RS) and “normal” rolls (RN).

The relation between λ and h is shown in Fig. 8a. As also found by other authors (e.g., Kelly 1984; Miura 1986), rolls occur at small λ and h, predominantly with λ < 5 km and h < 1 km. In our cases (with one exception discussed below), clear cellular patterns occurred with λ > 15 km and h > 1.4 km. The aspect ratio λ/h in the roll cases was between 2.6 and 6.5 and thus clearly smaller than in the cell cases, which had values up to λ/h = 12.6. On all seven days when crosswind vertical sections were flown at two different Δx, the downwind increase in h is accompanied by an increase in λ in such a way that λ/h increases, too.

This is more evident from Fig. 8b where the aspect ratio λ/h is displayed as function of the downwind distance Δx from the ice edge. However, there is a large variability in the λ/hx) relation from case to case both in the level of the λ/h values and in the rate it increases with Δx. The variability of λ/h appears to be larger in the region of the cellular patterns than in the roll region. Regarding all 20 cases as independent measurements the best logarithmic fit of λ/h for Δx between 25 and 1200 km is given by λ/h = 5.5 + 1.7 lnΔx, where Δx is taken in units of 100 km.

In Figs. 8a,b, a somewhat outstanding role is played by the two observations on 7 March 1991. On this day cells were observed having typical h values but untypically short wavelengths λ so that very small λ/h values result. The reason for this is not clear. This day was the only case when cells were not observed in an off-ice airflow but to the north of the Lofoten Islands in an easterly airflow coming over the mountains of northern Norway. In this case lee effects were involved. Cloud development started at a distance of about 150 km off the coast whereas in all other cases it started right at the ice edge. Table 1 shows that the cloud depth Δhc on 7 March 1991 at the first crosswind vertical section at Δx = 200 km is only 315 m and thus is rather small in comparison to the depth (h = 1115 m) of the boundary layer and also in comparison to the other cases. The weak cloud activity as a consequence of the lee effect may be one reason for the small aspect ratios on this day, which amounted to only λ/h = 3.0 at Δx = 200 km and λ/h = 3.7 at Δx = 450 km. This conclusion is based on model simulations of convection in cold-air outbreaks performed by Müller and Chlond (1996). With a convection-resolving large eddy simulation model they showed that a downwind increase of λ/h can only be achieved if sufficient cloud development is present. Without clouds the convective patterns retain a small aspect ratio around λ/h = 3.0 similar to that known from the linear Rayleigh instability theory (λ/h = 2.8).

The increasing importance of clouds in downwind direction from the roll pattern to the cellular pattern region is also indicated in Table 1. Cloud depth in the roll region is between 200 and 700 m, whereas in the cell region it is between 700 and 2000 m. Although there is a gradual decrease of cloud coverage N in downwind direction, the average cloud layer thickness if flattened homogeneously over the horizontal area, that is, ΔhcN, is nearly twice as large in the cell (660 m) than in the roll region (250–340 m). However, the relative area Fcl,rel of a vertical y–z cross section through the secondary flow pattern that is filled with clouds, that is,
i1520-0469-56-15-2613-e2
decreases from about 0.5 to 0.3 from the roll to the cell region. Thus, taken together, when moving in downwind direction the depth of the layer filled by clouds increases absolutely, but decreases in relation to the depth of the layer involved in the secondary flow.

b. Amplitudes of the secondary flow

Figure 9 represents a sketch of the secondary flow in the vertical y–z plane perpendicular to the mean wind. The sketch illustrates that the largest variations of the secondary flow variables in the crosswind y direction are to be expected at low levels for the lateral (υR) and longitudinal (uR) wind component as well as for temperature (θR) and water vapor mixing ratio mR. The largest variations of the vertical wind component wR are expected to occur at levels near the middle of the boundary layer. The latter holds also for the vertical fluxes: wRθR, wRmR, wRuR, and wRυR.

The amplitudes Aa (a = θR, mR, uR, υR, wR) of the secondary flow variables calculated according to (1) and classified with respect to the convective pattern according to Table 1 are listed in Table 2. They refer to the lowest permitted flight level of 90 m for θR, mR, and uR, υR. The amplitudes of wR refer to the level z(wmax) where the maximum of the secondary flow–scale vertical wind variance w2R was measured. Unfortunately, in many cases—especially in the roll cases with a very shallow boundary layer—there were only flights near the bottom and top of the boundary layer (see Fig. 3) so that the real maximum of AwR was not measured. In those cases the “90-m” values are quoted in brackets in Table 2.

Table 2 shows that the secondary flow temperature amplitude AθR is only a few tenths of a degree. The same order of magnitude holds for the water vapor amplitude AmR. However, AmR exhibits a clear increase in downwind direction from the roll to the cell patterns. The ratio of the heat content involved with temperature (AθR) and moisture fluctuations (AmR), that is, (L/cp)(AmR/AθR), where L is the latent heat of vaporation and cp the specific heat at constant pressure, increases from 0.65 for small rolls to 2.10 for “cells.” This underlines the importance of moisture processes as was derived by Brümmer (1997) from boundary layer heat budget calculations for the same cold-air outbreaks as discussed in this paper. He found that in the cell pattern region the latent heat release by condensation in clouds is the dominating term in the heat budget equation. The heating of the boundary layer by the surface sensible heat flux plays only a secondary role in this region but is the primary heat source in the roll pattern region.

The secondary flow wind amplitudes AuR, AυR, and AwR are on the order of 0.5–3 m s−1. All three amplitudes increase in a downwind direction and are roughly twice as large in cell patterns as in roll patterns. In general, the lateral wind component exhibits the largest amplitude of all three wind components and the vertical wind component the smallest one. However, there is a clear downwind trend concerning the ratio AuR/AυR. In accordance with the geometrical planform of the convective pattern, the ratio is about 0.66 in roll circulations and increases to about 1.0 in cell circulations, where there is no preferred space direction.

c. Total and secondary flow-scale fluxes

In this section the vertical fluxes of heat, moisture, and momentum by the secondary flow wRaR and by all scales of motion wa (a = θ, m, u, υ) are investigated with respect to the pattern form of the secondary flow and to their height dependence. The flux profiles are presented in Figs. 10a–d in a normalized form. The normalized height zN is calculated using cloud-base height hcb and cloud layer depth Δhc in the following way:
i1520-0469-56-15-2613-e3a
In this way the normalized height zN is between 0 ⩽ zN ⩽ 2. Here zN = 1 represents the cloud base and zN = 2 the top of the boundary layer. However, in order to take into account the extreme differences in hcb and Δhc between the roll and cell pattern regions, the height is again scaled by the mean values cb and Δc for the respective category of the secondary flow pattern, namely, the classes RS, RN, R/C, and C. Thus, the normalized but rescaled height zsc is defined as
i1520-0469-56-15-2613-e4a
In an analogous way, the flux at any height is first normalized by the total surface flux (wa)o,
i1520-0469-56-15-2613-e5a
and is then, in order to retain the basic flux differences between the different secondary flow patterns, scaled by the respective mean surface flux (denoted by a tilde) for each convective pattern class, that is, RS, RN, R/C, and C, as follows:
i1520-0469-56-15-2613-e6a
The values used to normalize and rescale the fluxes are listed in Table 3.
Since the surface fluxes (wa)o were not measured by the aircraft at the usual reference height of 10 m, the measured fluxes at 90 m are used instead. This is relatively uncritical in the case of the temperature flux (wθ)o and the moisture flux (wm)o, although they are slightly underestimated by this assumption, especially in the case of wθo (by about 10 %). However, in the case of the momentum flux, the measured flux at 90-m height could not by used instead of the surface momentum flux (wu)o. As defined (wu)o is negative, that is, u momentum is always transferred from the air to the sea surface. However, in 6 of the 20 cases a positive momentum flux was measured at 90-m height and even a linear downward extrapolation to 10-m height resulted in a positive momentum flux due to the few flight levels in the vertical and to the nonlinear shape of the momentum flux profile. In all of these abnormal cases the u-wind component decreased with height above 90 m. In these extremely unstable conditions the flux at 90-m height is dominated by the convective momentum flux, which composes the entire boundary layer and reacts to the momentum difference over a deep layer rather than by the turbulent flux that dominates at lower levels. Of course, these physical conditions also hold for the other 14 (normal) cases but did not lead to a change of the sign of the momentum flux at 90 m because the u component in these cases increased with height in the boundary layer, Thus, in order to normalize the momentum flux the surface value (wu)o was calculated from the bulk aerodynamic formula
uwocDU290
using the wind speed U90 measured at 90-m height and for simplicity a constant drag coefficient cD of 1.3 × 10−3, which is well within the range of neutral cD values given in Chou (1993). Unstable stratification and 90-m height (instead of 10 m) are opposing effects and thus will not alter cD too much from the chosen value.

To normalize the sensible and latent heat flux with the corresponding surface value is a meaningful and usual procedure, because the surface flux is the largest value and the decisive parameter in thermal convection. However, to normalize the momentum flux with its surface value may be questionable, because it is, in contrast to the heat and moisture flux, not always the maximum flux value and not the only decisive parameter. The momentum flux profile can depend on many factors, for example, the thermal wind.

Figures 10a–d show the total wa and the secondary flow–scale vertical fluxes wRaR of the variables a = θ, m, u, and υ. The results are grouped with respect to the four classes of convective patterns: RS, RN, R/C, and C. The curves within each class show some degree of variability from day to day depending on the actual conditions of air–sea temperature difference, wind speed, etc. The height variations within each curve represent systematic differences with height as well as sampling uncertainties resulting, for example, from a somewhat unrepresentative flight leg. In spite of possible uncertainties in the individual curves, the scatter of the curves within each category in Figs. 10a–d shows that the general differences between the secondary flow and the total flow as well as the general differences between the individual categories are significant. Only these general differences will be discussed in the following.

The sensible heat flux at the surface, (wθ)o, decreases in the downwind direction. The contribution of the secondary flow to the total flux increases from zero (as defined) at the surface to a maximum value in the subcloud layer and decreases more or less monotonously above. Thus, the secondary flow exports sensible heat from the lowest part of the boundary layer to the upper parts. This effect of the secondary flow is especially evident in the moisture flux profiles. The secondary flow moisture flux is zero at the base and the top of the boundary layer and has a broad maximum around the height range hcb < z < h/2 and, thus, moisture is transported by secondary flow motions from the subcloud to the cloud layer. The maximum latent heat flux by the secondary flow, (wRmR)max, is between 0.015 and 0.030 g kg−1 m s−1 in the four categories. Compared to the total surface latent heat flux, (wm)o, which is between 0.050 and 0.065 g kg−1 m s−1 in the four categories, the following relation results on the average:
wRmRmaxwmo
From Fig. 10b it appears that the ratio (wRmR)max/(wm)o increases from the roll to the cell region. However, there is a large scatter from case to case and the number of observations are too small to make a significant statement in this respect.

To point out the systematic differences between the total and secondary flow fluxes and the differences between the individual categories more clearly, the average sensible and latent heat flux profiles are displayed in Fig. 11. The total latent heat flux, mw, shows, concerning its profile shape, a systematic variation from the roll pattern region to the cell pattern region. Whereas the flux profile between the maximum surface value and the nearly vanishing top value is only slightly curved in the roll region, where a shallow cloud layer is present, it is strongly curved and even exhibits a maximum that is larger than the surface flux in the cell region, where deeper clouds are present.

Similar systematic variations of the moisture flux profiles were found in roll convection by Brümmer (1985) and in cell convection by Brümmer et al. (1986). Since the divergent subcloud moisture flux in the cell region would finally dry out the subcloud layer (what is not observed), there must be processes acting in the opposite direction. Possible candidates are a large-scale horizontal moisture convergence or the evaporation of precipitation. Assuming precipitation rates of about 4 mm day−1 in the cell convection region, as was estimated by budget analyses by Brümmer (1997) for the same cold-air outbreaks, evaporation rates needed to balance the moisture budget of the subcloud layer would be far beyond plausibility. Thus the large-scale moisture convergence remains as a possible balance process. This conjecture is supported (a) by the presence of large-scale convergence, which was analyzed from the mass budget estimates in Brümmer (1997); and (b) by a systematic downwind transition from an anticyclonically curved flow at distances Δx < 100 km from the ice edge to a cyclonically curved flow at distances Δx > 200 km, which was observed in the ARKTIS 1993 cold-air outbreaks (Brümmer 1996).

Furthermore, similar conditions are observed in situations of deep convection in the Tropics. Brümmer (1979) found that the vertical moisture flux in the subcloud layer was divergent (convergent) under disturbed (undisturbed) weather conditions with large-scale horizontal flow convergence (divergence) and deep (shallow) convection over the tropical Atlantic Ocean during the 1974 Global Atmospheric Research Program Atlantic Tropical Experiment. The relation between large-scale horizontal moisture convergence and deep convection is a repeatedly observed fact. It is also taken into account in the parameterization schemes for deep convection (e.g., Tiedtke 1989).

The systematic changes of the moisture flux profiles from the roll region to the cell region and the above-mentioned facts on the large-scale moisture convergence suggest the following hypothesis: a necessary condition for the transition from roll convection to mesoscale cell convection is the formation of deep clouds as a result of large-scale moisture convergence. This hypothesis is further supported in section 5a where it is found that the ratio of the condensational heating in clouds to the surface heating increases from less than 1 in the roll regions to more than 1 in the cell regions.

We return to Fig. 10 and the discussion of the fluxes of longitudinal (wu) and lateral momentum (wυ). The momentum flux profiles vary considerably from case to case within each category and a systematic trend from the roll region to the cell region is not present. This is a consequence of the different shapes of the wind profiles of both the longitudinal and the lateral wind component. Thus, the average flux profiles are not of general meaning. However, independent of the category of the convective pattern, the total momentum flux profiles exhibit the typical known features.

  1. The total wu flux decreases with height between the surface and at least cloud base, where the flux may be negative or positive. This is equivalent to an export of momentum and thus to a reduction of momentum in this layer, which of course has to be compensated by the mean pressure gradient force. These average conditions are in agreement with a profile where the u-wind component increases with height to a maximum value at or below cloud base.

  2. The total wυ flux is around zero at the surface and has more frequently positive values than negative ones at the levels above. Positive wυ fluxes suggest decreasing υ values with height. This is equivalent to a turning of the wind direction with height to the right as it is the typical case under the action of surface friction. Continuing this interpretation, negative wυ fluxes would suggest increasing υ values with height and a backing of wind direction. This occurs in situations when the cold-air advection is so strong that the thermal wind effect overcompensates the frictional wind veering.

As for the total momentum fluxes, also the secondary flow momentum flux profiles are highly variable and a systematic dependency neither with height nor from category to category is apparent. However, only with a very few exceptions, the secondary flow momentum flux has the same sign as the total flux and is smaller than the total flux. The ratio of secondary flow momentum flux to total momentum flux typically ranges between 0.1 and 0.7.

5. Characteristics of the basic flow in areas ofroll and cell convection

In this section the basic flow characteristics (temperature profile, wind profile) in the regions of the four convection classes (RS, RN, R/C, and C) are presented. The discussion is subdivided with respect to 1) the thermodynamic features and those parameters that affect thermal instability and aspect ratio and 2) the kinematic features and those parameters that affect dynamical instability and geometrical form of the secondary flow (linear or cellular). Finally, in section 7c an attempt is made to relate the magnitude of the secondary flow kinetic energy to basic flow parameters.

a. Thermodynamic features of the basic flow

Table 4 summarizes the thermodynamic conditions in the surface, boundary, cloud, and inversion layers in the regions of RS, RN, R/C, and C. In the surface layer (termed here as the layer between the sea surface and the lowest flight level of 90-m height), the stratification is rather unstable. The air–sea temperature difference Δθas gradually changes from extremely unstable values in the area of small rolls near the ice edge to moderately unstable values in the area of cells. This goes parallel with the general downstream increase of the mean potential air temperature θ. Depending on the synoptic conditions, the wind speed at 90-m height U90 shows a large degree of variability from case to case in each convective category. Thus the mean wind differences from one convection category to the other cannot be regarded as systematic for the pattern transition. A similarly large degree of variability is present for the Monin–Obukhov length LM, which was calculated according to
i1520-0469-56-15-2613-e9
In (9) g is the acceleration of gravity, k = 0.4 is the von Kármán constant, and the bulk formulas are used for calculation of momentum and heat flux. For simplicity the corresponding transfer coefficients cD and cH are assumed to be equal as cD = cH = 1.3 × 10−3. This assumption does not influence the results for LM significantly. In all convection categories LM is clearly unstable but shows no clear trend from the roll to the cell region. The same is true for h/LM, regarded as a stability parameter for the entire boundary layer. Thus, both LM and h/LM do not appear to be useful measures in distinguishing between roll and cell convection.

The stratification in the overlying part of the boundary layer, between 90 m and h, is slightly stable as represented by the corresponding temperature difference Δθbl. On the average, the stability Δθbl/(h − 90 m) is about twice as large in the cell region as in the roll region.

The thermal stability conditions in the boundary layer are usually described by the Rayleigh number Ra. It represents the ratio between viscous and buoyancy force and is determined here from the relation
i1520-0469-56-15-2613-e10
where KM and KH are the turbulent diffusion coeficients for momentum and heat, respectively. Since KM cannot be measured directly, it was calculated according to Louis (1979) as
i1520-0469-56-15-2613-e11
for the surface layer with depth Δz = 90 m. In (10) the mixing length, l, is given by the relation
i1520-0469-56-15-2613-e12
with z = 45 m and λ = 100 m; the Richardson number Ri in its bulk version as
i1520-0469-56-15-2613-e13
assuming Δv = v90 − 0; and the stability function F(Ri) is defined as
i1520-0469-56-15-2613-e14
with the coefficient b = 9.4 and c = 7.4bl2[(z + Δz/z)1/3 − 1]3/2/(z1/2Δz3/2). Because of the uncertainties that are inherent in an approximation of KM like in (11), and for the sake of simplicity, we assume that KM = KH in (10) and that the surface layer values of KM are roughly representative of the average values in the boundary layer.

The values of KM estimated in this way are listed in Table 4 and vary between 29 and 71 m2 s−1. They differ only by a factor of about 2 so that a constant KM, as it is often assumed (e.g., Walter 1986) for the calculation of Ra, is not too much of a simplification. Our Ra values increase in downstream direction from the roll to the cell region. This holds not only for the category averages, but also for each of those days when measurements at two distances Δx from the ice edge were made. The Rayleigh number Ra is about 10 to 20 times larger in the cell area than in the roll area. It appears from Table 4 that the transition from roll convection to cell convection occurs in the range between 0.8 × 106 < Ra < 1.0 × 106.

Walter (1986) also investigated the relation between Ra and convection type. He scaled Ra with the critical Rayleigh number Rac from the linear theory (Rac = 1708) and observed rolls for 20 ⩽ Ra/Rac ⩽ 100 and cells for Ra/Rac ≥ 250. In 15 out of 16 cases in the convection categories RS, RN, and C, our values of Ra/Rac do not contradict Walter’s classification. However, the range of Ra/Rac values in our transitional category R/C is much broader (88 ⩽ Ra/Rac ⩽ 533) than that given by Walter (100 < Ra/Rac < 250).

Beside surface heating in promoting thermal instability, the latent heat release by condensation in clouds is another important process. In connection with boundary layer moisture and heat budget calculations for cold-air outbreaks from the Arctic sea ice, Brümmer (1997) estimated the net condensation/evaporation, ce, in clouds. For those cases where the budget-study days are identical to the days investigated here, the ce values are listed in Table 4. The heating effect of net condensation is converted in Table 4 to an equivalent heat flux, (L/cp)h(ce), at the bottom of the boundary layer and can thus be directly compared with the surface buoyancy flux (θυw)90. Table 4 shows that the condensational heating increases gradually by a factor of 2 to 3 from the roll to the cell convection region and that it is as large or even larger than the surface buoyancy flux in the cell region. It is thus hypothezised that condensation is a necessary process for the generation of cell convection and that its heating effect must at least reach the magnitude of the surface heating.

Numerical simulations by Müller (1995) show that the condensation process is a necessary condition for the transition from convection with small aspect ratios to convection with large aspect ratios as it is observed in mesoscale cellular convection. Model results of Rao and Agee (1996) show that even the mode of precipitation (water or ice) and the associated microphysical processes have a significant effect on the structure of the convective boundary layer. A broadening of the aspect ratio is also known to be caused by increasing stability in the cloud layer (e.g., Bjerknes 1938; Chlond 1988). In agreement with this idea are the above-mentioned increase of Δθbl (Table 4) and the corresponding increase of aspect ratio λ/h (Table 1). Except for 8 March 1991, this relation holds on all days when measurements at two distances Δx were made (11, 19, 20, 24, and 25 March 1993; 7 March 1991). Since both condensation and stability are working or are present simultaneously, their individual contribution to the observed aspect ratio broadening cannot be seperated. Furthermore, condensation and stability are interacting in a sense that increasing stability reduces condensation and vice versa.

The stability in the inversion topping the cloud layer decreases systematically from the roll to the cell region as it is represented by the potential temperature difference Δθ300ct over a 300-m-deep layer above cloud top (Table 4). This is a consequence of the continuous warming of the boundary layer air over the open water, while the air above the boundary layer remains nearly unaffected. The weaker stability of the inversion in the cell region allows more variability in the heights of the cloud tops than in the roll region. One exception was on 25 February 1991 when closed cellular convection and a large value of Δθ300ct were observed in a situation of high pressure subsidence with flat cloud tops and high cloud coverage N (see Table 1). All other cell cases presented here refer to open cellular convection.

b. Kinematic features of the basic flow

The vertical profile of the horizontal wind plays an important role for the organization of the convection. Wind shear (e.g., Asai 1970), curvature (e.g., Kuettner 1971), and inflection point (e.g., Lilly 1966) may cause longitudinal organization. In case of thermal instability with wind shear and curvature the convection is organized in rolls that are parallel to the shear vector or to the vertical plane in which the curvature is present. In case of inflection point instability the roll axes are oriented perpendicular to the plane in which the inflection point is observed.

In order to characterize the vertical wind structure in a simple (but sufficient) manner the wind vector has been taken at four levels only: 90 m, h/2, h, and h + 300 m. The resulting wind profiles relative to the corresponding wind at 90 m in each convection case are shown in Fig. 12 while Fig. 13 shows the mean wind values for each convection category in absolute wind units. Based on the wind at these levels, the layer-averaged boundary layer wind, the shear across the boundary layer and across the inversion layer, as well as the curvature in the boundary layer were calculated and the results are listed in Table 5. The curvature was calculated according to
i1520-0469-56-15-2613-e15
The layer-averaged boundary layer wind does not appear to be useful measure to distinguish between rolls and cells. It varies from small to large values within each category and simply reflects the various gradients of the large-scale pressure field. The wind shear, however, is significantly smaller in the cell than in the roll regions. Taking the category averages of the shear magnitude the difference between rolls and cells is on the order of a factor of 2–3. This holds for the shear in the boundary layer as well as in the inversion layer. Independent of the convective pattern form, the shear in the inversion layer is always larger than that in the boundary layer, which reflects the different turbulence states within both layers.

Assuming 10−5 m−1 s−1 as a critical magnitude of the curvature above which convection is organized in rolls (Kuettner 1971), the curvature ∂2u/∂z2 in the x direction along the mean wind is larger than the critical one in nearly all roll convection cases. This is in accordance with the observation that the cloud streets were aligned in the downwind direction. On the other hand the curvature in all cell cases is clearly below the critical value. With only one exception (11 March 1993 in the RS category) our observations support the critical value of the wind curvature that was suggested by Kuettner (1971) and was also based on observational evidence.

In several cases of roll convection the author (as aircraft scientist on board the research aircraft Falcon) observed by eye periodic variations of the cloud field in the along-roll direction giving the impression of a second instability mechanism overlying more or less perpendicularly the primary rolls (Fig. 14). It is supposed that the overlying variations originate from the inflection point instability mechanism operating additionally at levels near the top of the boundary layer. In Table 5 the occurrence of an inflection point in the wind profile is indicated by its height hip relative to h (two values for the same case refer to the two profiles at both sides of the vertical plane in which the horizontal flight legs were flown; see Fig. 3). In the roll categories RS and RN, an inflection point in the u component occurs in 75% of the cases; it is situated relatively close above the top of the boundary layer. On the other hand, an inflection point in the υ component is less frequent and situated at a much higher level above the top. These figures at least do not contradict the presumption on the origin of the secondary periodic cloud variations. In the transition region R/C and the cell region C an inflection point is less frequent but occurs in both wind components and relatively close above h. In this case secondary periodic variations may have also occurred but are more difficult to be recognized by eye in the cloud field because of the underlying cellular pattern, the larger wavelength of the cell pattern, and the smaller cloud coverage as compared to an underlying roll pattern.

c. Relation between the basic flow and the kinetic energy of the secondary flow

One aim of the investigations on the secondary flow, besides the study of its properties, is to relate these properties to parameters of the basic flow, in order to account for the secondary flow effects in large-scale weather and climate models that are not able to resolve the secondary flow explicitly. Since in this study parameters of both the basic and the secondary flow have been measured or derived from the measurements, a first attempt is made to relate the layer-integrated secondary flow–scale kinetic energy,
i1520-0469-56-15-2613-e16
to the Rayleigh number Ra, regarded as a measure of the stability conditions of the basic flow in the boundary layer. The secondary flow–scale kinetic energy as well as the individual contributions by the three wind components are listed in Table 6 and the relation between Ekin,R and Ra (taken from Table 4) is presented in Fig. 15.
Figure 15 shows a general increase of Ekin,R with increasing Ra. Such an increase also holds for five of those seven cases when two vertical cross sections at two different distances Δx from the ice edge were flown on the same day (11 March 1993, 24 March 1993, 25 March 1993, 7 March 1991, 8 March 1991). These days are connected by a dashed line in Fig. 15. On the other two days (19 March 1993, 20 March 1993) this increase does not hold and instead a slight decrease of Ekin,R with Ra is found. The best fitting power law to the Ekin,R (Ra) relationship is given by
Ekin,R0.4046
According to Fig. 15, Ekin,R does not appear to increase at infinity with increasing Ra but appears to level off and to approach a kind of “saturation” value. Although Ekin,R increases clearly with Ra it can be seen from Table 6 that the kinetic energy density (kinetic energy per unit height interval) ½v2R does not increase significantly with Ra. Thus, it may be concluded from the observations that with increasing instability of the basic flow (increasing Ra) the secondary flow does not circulate at a higher average speed but does only affect a deeper layer at more or less the same circulation speed.

6. Concluding remarks

Most papers on organized convection deal either with the geometric dimensions of the secondary flow or with the parameters of the basic flow. In this study aircraft measurements conducted in 13 cold-air outbreaks from the Arctic sea ice over the open water of the Greenland and Barents Seas could be used, which allowed the determination of both the secondary flow characteristics (geometric dimensions, variances, and vertical fluxes) and the basic flow characteristics (thermodynamic and kinematic state). The data are a rather unique set compared to others discussed in the reviewed literature.

The measurements were subdivided into four classes with respect to the geometrical planform of the organized convection: longitudinal rolls with relatively small or with larger horizontal wavelengths, transitional forms between rolls and cells, and cells. The classification as made in Table 1, of course, bears some degree of arbitrariness and could have been made somewhat differently. For example, the first part of 11 March 1993 could have been put from the category RS to the category RN (it was not because of the very close distance Δx from the ice edge). Or it may be discussed whether 7 March 1991, the day with the very small cells, should have been put together with the much broader cells into the same cell category; or whether 25 February 1991, the day with the closed cells, belongs in the same cell category as the open cell cases. In spite of this arbitrariness, the intention of this paper, to bring a systematic order into the measurements made at different distances Δx from the ice edge and within different convective organization forms, was successful. Many of the results presented above show a more or less systematic variation in downstream direction from the roll to the cell convection region.

Due to the limited flight duration of about 3.5 h (corresponding to a distance of about 1300 km including the transit between the airfield and the operational area) a compromise had to be made during the cold-air outbreak aircraft missions between the demand to have a narrow vertical coverage (many flight levels above each other) on the one hand and the demand to have a wide horizontal coverage (airmass transformation over long distances Δx) on the other hand. As a consequence of this compromise, all flight sections in the category of the small rolls contain only two flight levels, namely, at 90 m and near cloud top. This turned out to be a real disadvantage in this paper. At least one additional flight at cloud base or near the middle of the boundary layer would have been helpful to estimate the maximum of the wR variance or of the mRwR flux of the secondary flow.

A first attempt has been made in this paper to relate a secondary flow variable (here the kinetic energy Ekin,R) to a parameter of the basic flow that represents the primary driving instability mechanism (here the thermal instability represented by the atmospheric Rayleigh number Ra). Figure 15 is regarded as a first quantitative hint with respect to an interrelationship between Ekin,R and Ra. However, it contains a large degree of scatter that is due to uncertainties both in Ekin,R and Ra. The most uncertainty in Ekin,R [defined as a vertical integral;see Eq. (16)] arises from the poor vertical coverage but may be overcome by a better vertical average. The most uncertainty in Ra arises from the uncertainty of the diffusion coefficents KM and KH. The resulting uncertainty in Ra is supposed to be at least on the order of a factor of 2 and there is no obvious way to improve this situation.

In order to conclude from the basic flow state to the conditions of the secondary flow the knowledge of one basic flow parameter (e.g., Ra) is not sufficient. At least one additional flow parameter characterizing the kinematic basic state, for example, a shear Reynolds number, which may be defined as
i1520-0469-56-15-2613-e18
where Δu is the wind difference over the depth of the boundary layer h, is necessary to judge the planform of the secondary flow. Furthermore, it is known that the stability of the capping inversion above the boundary layer plays an important role for the pattern form and can determine whether open or closed cells will develop. Thus, for example, a second Rayleigh number for the layer above the boundary layer is helpful. In addition, the results in Table 4 suggest that condensation heating in relation to surface heating is an important process for the generation of cell convection. Thus, the ratio of condensation heating to surface heating (L/cp)h(ce)/(wθυ)o appears to be another relevant number to describe the convective conditions in the atmosphere compared to those in the laboratory.

As a next step, the above-mentioned attempt to relate secondary flow variables to parameters of the basic flow should be expanded, for example, to find a relation between the secondary flow–scale fluxes and Ra. At least for the secondary flow–scale moisture flux mRwR, which exhibits a rather clear vertical profile shape (Fig. 11) and which may be characterized simply by the maximum value (mRwR)max, our measurements suggest that it should be possible to find such a relationship. To find a corresponding relation for the temperature and momentum fluxes appears to be much more difficult because of the more complex profile shape of these fluxes.

This paper is regarded as the third of a trilogy on cold-air outbreaks from the Arctic sea ice. The two preceeding papers are Brümmer (1996), which deals primarily with the bulk modification of the boundary layer in downstream direction, and Brümmer (1997), which deals with the mass and energy budgets at different distances from the ice edge.

Acknowledgments

This paper was sponsored by the German Science Foundation under Sonderforschungsbereich 318 “Klimarelevante Prozesse im System Ozean, Atmosphäre, Kryosphäre.” I want to thank B. Busack and B. Löbe for processing the aircraft data and calculating the secondary flow–scale fluxes and variances, S. Thiemann and M. Lüdicke for preparing the figures, and S. Lehmann for typing the manuscript.

REFERENCES

  • Agee, E. M., 1984: Observations from space and thermal convection—A historical perspective. Bull. Amer. Meteor. Soc.,65, 938–949.

  • ——, 1987: Mesoscale cellular convection over the oceans. Dyn. Atmos. Oceans,10, 317–341.

  • Asai, T., 1970: Stability of a plane parallel flow with variable vertical shear and unstable stratification. J. Meteor. Soc. Japan,48, 129–139.

  • Atkinson, B. W., and J. W. Zhang, 1996: Mesoscale shallow convection in the atmosphere. Rev. Geophys.,34, 403–431.

  • Bjerknes, J., 1938: Saturated ascent of air through a dry-adiabatically descending environment. Quart. J. Roy. Meteor. Soc.,64, 325–330.

  • Brown, R. A., 1970: A secondary flow model for the planetary boundary layer. J. Atmos. Sci.,27, 742–757.

  • ——, 1972: The inflection point instability problem for stratified boundary layers. J. Atmos. Sci.,29, 850–859.

  • ——, 1980: Longitudinal instabilities and secondary flows in the planetary boundary layer: A review. Rev. Geophys.,18, 683–697.

  • ——, and W. T. Liu, 1982: An operational large-scale marine planetary boundary layer model. J. Appl. Meteor.,21, 261–269.

  • ——, and R. Foster, 1994: On PBL models for general circulation models. Global Atmos.–Ocean Syst.,2, 163–183.

  • Brümmer, B., 1979: Boundary layer processes during the life cycle of GATE convective disturbances. Contrib. Atmos. Phys.,52, 306–330.

  • ——, 1985: Structure, dynamics and energetics of boundary layer rolls from KonTur aircraft observations. Contrib. Atmos. Phys.,58, 237–254.

  • ——, Ed., 1992: ARKTIS 1991: Report on the field phase with examples of measurements. Berichte aus dem Zentrum für Meeres- und Klimaforschung, Hamburg, Reihe A, No. 3, 216 pp. [Available from Meteorologisches Institut, Bundesstr. 55, Hamburg D-20146, Germany.].

  • ——, Ed., 1993: ARKTIS 1993: Report on the field phase with examples of measurements. Berichte aus dem Zentrum für Meeres- und Klimaforschung, Hamburg, Reihe A, No. 11, 270 pp. [Available from Meteorologisches Institut, Bundesstr. 55, Hamburg D-20146, Germany.].

  • ——, 1996: Boundary layer modification in wintertime cold-air outbreaks from the Arctic sea ice. Bound.-Layer Meteor.,80, 109–125.

  • ——, 1997: Boundary layer mass, water, and heat budgets in wintertime cold-air outbreaks from the Arctic sea ice. Mon. Wea. Rev.,125, 1824–1837.

  • ——, T. Fischer, and S. Zank, 1986: Aircraft observations of open cellular structures during KonTur. Contrib. Atmos. Phys.,59, 162–184.

  • Chlond, A., 1988: Numerical and analytical studies of diabatic heating effect upon flatness of boundary layer rolls. Contrib. Atmos. Phys.,61, 312–329.

  • ——, 1992: A three-dimensional simulation of a cloud street development during a cold-air outbreak. Bound.-Layer Meteor.,58, 161–200.

  • Chou, S.-H., 1993: A comparison of airborne eddy correlation and bulk aerodynamic methods for ocean-air turbulent fluxes during cold-air outbreaks. Bound.-Layer Meteor.,64, 75–100.

  • ——, and J. Zimmerman, 1989: Bivariate conditional sampling of buoyancy flux during an intensive cold-air outbreak. Bound.-Layer Meteor.,46, 93–112.

  • ——, and M. P. Ferguson, 1991: Heat fluxes and roll circulations over the western Gulf Stream during an intense cold-air outbreak. Bound.-Layer Meteor.,55, 255–283.

  • Etling, D., 1971: The stability of an Ekman boundary layer flow as influenced by the thermal stratification. Contrib. Atmos. Phys.,44, 168–186.

  • ——, and R. A. Brown, 1993: Roll vortices in the planetary boundary layer: A review. Bound.-Layer Meteor.,65, 215–248.

  • Hein, P. F., and R. A. Brown, 1988: Observations of longitudinal roll vortices during Arctic cold air outbreaks over open water. Bound.-Layer Meteor.,45, 177–199.

  • Helfand, H. M., and E. Kalnay, 1983: A model to determine open or closed cellular convection. J. Atmos. Sci.,40, 631–650.

  • Kelly, R. D., 1984: Horizontal roll and boundary layer interrelationships observed over Lake Michigan. J. Atmos. Sci.,41, 1816–1826.

  • Kuettner, J., 1971: Cloud bands in the earth’s atmosphere: Observations and theory. Tellus,23, 404–425.

  • Lilly, D. H., 1966: On the instability of Ekman boundary layer flow. J. Atmos. Sci.,23, 481–494.

  • Louis, J. F., 1979: A parametric model of vertical eddy fluxes in the atmosphere. Bound.-Layer Meteor.,17, 187–202.

  • Mason, P. J., 1989: Large eddy simulation of the convective atmospheric boundary layer. J. Atmos. Sci.,46, 1492–1516.

  • Miura, Y., 1986: Aspect ratios of longitudinal rolls and convection cells observed during cold-air outbreaks. J. Atmos. Sci.,43, 26–39.

  • Müller, G., 1995: Mesoskalige Zellularkonvektion in Abhängigkeit von unterschiedlichen physikalischen Prozessen und synoptischen Randbedingungen: Numerische simulationen. Ph.D. dissertation, Berichte aus dem Zentrum für Meeres- und Klimaforschung, Hamburg, No. 16, 144 pp. [Available from Meteorologisches Institut, Bundesstr. 55, Hamburg D-20146, Germany.].

  • ——, and A. Chlond, 1996: Three-dimensional numerical study of cell broadening during cold-air outbreaks. Bound.-Layer Meteor.,81, 289–323.

  • Raasch, S., 1990: Numerical simulation of the development of the convective boundary layer during a cold-air outbreak. Bound.-Layer Meteor.,52, 349–375.

  • Rao, G.-S., and E. M. Agee, 1996: Large eddy simulation of turbulent flow in a marine convective boundary layer with snow. J. Atmos. Sci.,53, 86–100.

  • Rosmond, T. E., 1973: Mesoscale cellular convection. J. Atmos. Sci.,30, 1392–1409.

  • Tiedtke, M., 1989: A comprehensive mass flux scheme for cumulus parameterization in large-scale models. Mon. Wea. Rev.,117, 1779–1800.

  • Walter, B. A., 1986: The mesoscale organization, dynamics, and evolution of the marine planetary boundary layer during cold-air outbreaks. Ph.D. dissertation, University of Washington, 200 pp. [Available from University Microfilms, 300 North Zeeb Rd., Ann Arbor, MI 48106.].

  • Wayland, R. J., and S. Raman, 1989: Mean and turbulent structures of a baroclinic marine boundary layer during the 28 January 1986 cold-air outbreak (GALE 86). Bound.-Layer Meteor.,48, 227–254.

Fig. 1.
Fig. 1.

National Oceanic and Atmospheric Administration infrared satellite imagery at 1249 UTC 25 March 1993, showing a cold-air outbreak off the ice edge near Spitsbergen. Linear cloud patterns change to cellular cloud structures with increasing distance from the ice edge. The flight pattern on this day is marked by white lines.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 2.
Fig. 2.

Geographical location of the field experiments ARKTIS 1991 and ARKTIS 1993.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 3.
Fig. 3.

Three-dimensional display of a typical flight pattern of the two research aircraft Falcon-20 and Dornier-128 during aircraft missions in cold-air outbreaks during the field experiments ARKTIS 1991 and ARKTIS 1993.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 4.
Fig. 4.

Vertical profiles of potential temperature θ, specific humidity q, and wind components u and υ parallel and normal to mean wind direction (DD = 10°) measured at 1209 UTC 24 March 1993 at 78.4°N and 6.1°E at a distance of 160 km from the ice edge. The relevant basic flow parameters are indicated: boundary layer top h; cloud base hcb; air–sea differences Δθas, Δqas; vertical differences across the boundary layer Δθbl, Δqbl, Δubl, Δυbl and between h and h + 300 m, Δθct, Δqct, Δuct, Δυct, as well as of inflection point height hip in the u-wind profile.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 5.
Fig. 5.

Time series of downwelling longwave radiation L↓, temperature T, water vapor mixing ratio m, and longitudinal (u), lateral (υ), and vertical (w) wind component measured during a crosswind flight leg at 92 m on 24 March 1993 within a field of roll convection. The centers of the overlying cloud streets are marked by thin vertical lines. Time marks are every 10 s corresponding to a distance of 1 km.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 6.
Fig. 6.

Variance spectra SLL of longwave radiation weighted by frequency f and calculated from time series measured during four vertically staggered crosswind flight legs on 24 March 1993 within a field of roll convection. Frequencies f1 and f2 mark the limits of the roll-scale wavelength range. Frequency f can be converted to wavelength λ by λ = c/f with the aircraft speed of c = 100 m s−1.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 7.
Fig. 7.

Variance spectra Saa (a = u, υ, w, θ, m) weighted by frequency f and calculated from the time series shown in Fig. 5. Frequencies f1 and f2 mark the limits of the roll-scale wavelength range and are taken from Fig. 6.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 8.
Fig. 8.

(a) Wavelength λ vs height h of the secondary flow pattern and (b) aspect ratio λ/h vs ice edge distance Δx measured by aircraft in 20 cases during the field experiments ARKTIS 1991 and 1993. Encircled numbers mark the day of observation (see Table 1). Dashed lines connect cases observed at different Δx on the same day. Thin straight lines in (a) mark the λ/h bounds of roll patterns and cell patterns. The curved line in (b) represents the least squares logarithmic fit: λ/h = 5.5 + 1.7 lnΔx, where Δx is in units of 100 km.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 9.
Fig. 9.

Sketch of the secondary flow circulation in the crosswind vertical y–z plane. Flight levels at the lowest permitted height (90 m) and in the middle of the boundary layer are indicated by dashed lines.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 10.
Fig. 10.

Total and secondary flow–scale vertical fluxes of (a) sensible heat, wθ; (b) latent heat, mw; (c) longitudinal momentum, wu;and (d) lateral momentum, wυ. The fluxes are normalized as described in section 4c and are classified with respect to the convective pattern form in four categories: rolls (small), rolls (normal), rolls/cells, and cells. Numbers at the curves represent the day of observation (see Table 1). Thin horizontal lines mark cloud base and cloud top.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 11.
Fig. 11.

Total (thick lines) and secondary flow–scale (dashed lines) vertical fluxes of latent heat, wm (above), and sensible heat, wθ (below), in regions of different convective patterns: rolls (small), rolls (normal), rolls/cells, and cells. The curves represent averages of the corresponding curves in Fig. 10. Thin horizontal lines mark cloud base and cloud top.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 12.
Fig. 12.

Vertical profiles of longitudinal wind component u and lateral wind component υ relative to the corresponding values at 90-m height. Numbers indicate the day of observations (see Table 1). Horizontal lines represent cloud base and cloud top. The heights are scaled according to (4a), (4b).

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 13.
Fig. 13.

Vertical profiles of longitudinal and lateral wind components u and υ averaged for each convective category. Short horizontal lines represent standard deviation and long horizontal lines represent cloud base and cloud top. The heights are scaled according to (4a), (4b).

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 14.
Fig. 14.

Sketch of two superimposed cloud patterns observed during several flights within roll convection. The convective pattern along the x direction originates from thermal instability due to surface heating and is organized by wind shear and curvature of the u-wind component while the convective pattern along the y direction is supposed to originate from the inflection point instability due to an inflection point in the u-wind component close to the top of the boundary layer.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Fig. 15.
Fig. 15.

Secondary flow–scale kinetic energy ½v2Rh integrated over the depth h of the boundary layer vs Rayleigh number Ra. The numbers mark the day of observation (see Table 1). The full line represents the least squares power law fit Ekin,R = 3.69Ra0.40. The dashed lines connect the two observations of increasing Ekin,R with increasing Ra on those five of the seven days when observations were made at two different distances Δx along a streamline from the ice edge.

Citation: Journal of the Atmospheric Sciences 56, 15; 10.1175/1520-0469(1999)056<2613:RACCIW>2.0.CO;2

Table 1.

Height h, wavelength λ, aspect ratio λ/h of the secondary flow pattern and cloud base hcb, cloud layer depth Δhc, and cloud cover N of the corresponding cloud field measured by aircraft in 13 cold-air outbreaks during the field experiments ARKTIS 1991 and 1993. The day of observation, the sign of abbreviation (as used in Figs. 8, 10, 12, and 15) and the distance Δx of the observation from the ice edge are also given. The observations are classified with respect to the convective secondary flow pattern into four classes: small rolls (RS), normal rolls (RN), transitional forms between rolls and cells (R/C), and cells (C).

Table 1.
Table 2.

Amplitudes Aa (a = θ, m, u, υ, w) of the secondary flow variables classified with respect to the convective pattern. Values refer to the lowest permitted flight level at about 90 m except for Aw, which refers to the level z(wmax). The Aw and z(wmax) values in brackets indicate cases when Aw had to be taken from the 90-m level instead, from a flight level situated nearer to the middle of the boundary layer.

Table 2.
Table 3.

Total vertical fluxes of temperature (wθ)90, moisture (wm)90, and momentum at 90-m height used instead of the surface fluxes in order to normalize the flux profiles in Fig. 10. The bulk momentum flux cDU290 is used for normalization because the measured momentum flux at 90 m, (wu)90, is not representative for the surface flux. The table also lists characteristic velocity scales, that is, friction velocity u;zz, convective velocity w;zz, convective temperature scale θ;zz, and convective moisture scale m;zz.

Table 3.
Table 4.

Thermodynamic conditions in the surface, boundary, cloud, and inversion layers. Here Δθas is the potential temperature difference between 90 m and sea surface, U90 the 90-m wind magnitude, θ the mean potential temperature, LM the Monin–Obukhov length, h the boundary layer depth, Δθbl the temperature difference between the top of the boundary layer and 90 m, KM the turbulent diffusion coefficient for momentum, Ra the Rayleigh number, (L/cp)h(ce) an equivalent heat flux into the boundary layer due to latent heat release by net condensation/evaporation, (L/cp)h(ce)/(wθυ)90 the ratio of the heating due to net condensation/evaporation and due to surface flux, and Δθ300ct the temperature difference over a 300-m-deep layer above cloud top (h).

Table 4.
Table 5.

Dynamic conditions in and above the boundary layer. Here (ũ, υ̃) are the longitudinal and lateral wind components averaged over the depth h of the boundary layer; v90(u90, υ90), vh(uh, υh), and vh+300(uh+300, υh+300) are the wind vectors at 90 m, at the top, and 300 m above the top of the boundary layer, respectively; ∂2u/∂z2 and ∂2υ/∂z2 are the curvature in the boundary layer; huIP/h and hυIP/h are the normalized heights of the inflection point in the u- and υ-wind profiles at both sides of the crosswind horizontal flight legs (see Fig. 3). Percentage numbers in the “mean” rows give the frequency of the occurrence of an inflection point in the wind profiles.

Table 5.
Table 6.

Secondary flow–scale kinetic energy integrated over the depth h of the boundary layer and the contributions by the individual wind components. Tilde denotes averages over the depth of the boundary layer.

Table 6.
Save
  • Agee, E. M., 1984: Observations from space and thermal convection—A historical perspective. Bull. Amer. Meteor. Soc.,65, 938–949.

  • ——, 1987: Mesoscale cellular convection over the oceans. Dyn. Atmos. Oceans,10, 317–341.

  • Asai, T., 1970: Stability of a plane parallel flow with variable vertical shear and unstable stratification. J. Meteor. Soc. Japan,48, 129–139.

  • Atkinson, B. W., and J. W. Zhang, 1996: Mesoscale shallow convection in the atmosphere. Rev. Geophys.,34, 403–431.

  • Bjerknes, J., 1938: Saturated ascent of air through a dry-adiabatically descending environment. Quart. J. Roy. Meteor. Soc.,64, 325–330.

  • Brown, R. A., 1970: A secondary flow model for the planetary boundary layer. J. Atmos. Sci.,27, 742–757.

  • ——, 1972: The inflection point instability problem for stratified boundary layers. J. Atmos. Sci.,29, 850–859.

  • ——, 1980: Longitudinal instabilities and secondary flows in the planetary boundary layer: A review. Rev. Geophys.,18, 683–697.

  • ——, and W. T. Liu, 1982: An operational large-scale marine planetary boundary layer model. J. Appl. Meteor.,21, 261–269.

  • ——, and R. Foster, 1994: On PBL models for general circulation models. Global Atmos.–Ocean Syst.,2, 163–183.

  • Brümmer, B., 1979: Boundary layer processes during the life cycle of GATE convective disturbances. Contrib. Atmos. Phys.,52, 306–330.

  • ——, 1985: Structure, dynamics and energetics of boundary layer rolls from KonTur aircraft observations. Contrib. Atmos. Phys.,58, 237–254.

  • ——, Ed., 1992: ARKTIS 1991: Report on the field phase with examples of measurements. Berichte aus dem Zentrum für Meeres- und Klimaforschung, Hamburg, Reihe A, No. 3, 216 pp. [Available from Meteorologisches Institut, Bundesstr. 55, Hamburg D-20146, Germany.].

  • ——, Ed., 1993: ARKTIS 1993: Report on the field phase with examples of measurements. Berichte aus dem Zentrum für Meeres- und Klimaforschung, Hamburg, Reihe A, No. 11, 270 pp. [Available from Meteorologisches Institut, Bundesstr. 55, Hamburg D-20146, Germany.].

  • ——, 1996: Boundary layer modification in wintertime cold-air outbreaks from the Arctic sea ice. Bound.-Layer Meteor.,80, 109–125.

  • ——, 1997: Boundary layer mass, water, and heat budgets in wintertime cold-air outbreaks from the Arctic sea ice. Mon. Wea. Rev.,125, 1824–1837.

  • ——, T. Fischer, and S. Zank, 1986: Aircraft observations of open cellular structures during KonTur. Contrib. Atmos. Phys.,59, 162–184.

  • Chlond, A., 1988: Numerical and analytical studies of diabatic heating effect upon flatness of boundary layer rolls. Contrib. Atmos. Phys.,61, 312–329.

  • ——, 1992: A three-dimensional simulation of a cloud street development during a cold-air outbreak. Bound.-Layer Meteor.,58, 161–200.

  • Chou, S.-H., 1993: A comparison of airborne eddy correlation and bulk aerodynamic methods for ocean-air turbulent fluxes during cold-air outbreaks. Bound.-Layer Meteor.,64, 75–100.

  • ——, and J. Zimmerman, 1989: Bivariate conditional sampling of buoyancy flux during an intensive cold-air outbreak. Bound.-Layer Meteor.,46, 93–112.

  • ——, and M. P. Ferguson, 1991: Heat fluxes and roll circulations over the western Gulf Stream during an intense cold-air outbreak. Bound.-Layer Meteor.,55, 255–283.

  • Etling, D., 1971: The stability of an Ekman boundary layer flow as influenced by the thermal stratification. Contrib. Atmos. Phys.,44, 168–186.

  • ——, and R. A. Brown, 1993: Roll vortices in the planetary boundary layer: A review. Bound.-Layer Meteor.,65, 215–248.

  • Hein, P. F., and R. A. Brown, 1988: Observations of longitudinal roll vortices during Arctic cold air outbreaks over open water. Bound.-Layer Meteor.,45, 177–199.

  • Helfand, H. M., and E. Kalnay, 1983: A model to determine open or closed cellular convection. J. Atmos. Sci.,40, 631–650.

  • Kelly, R. D., 1984: Horizontal roll and boundary layer interrelationships observed over Lake Michigan. J. Atmos. Sci.,41, 1816–1826.

  • Kuettner, J., 1971: Cloud bands in the earth’s atmosphere: Observations and theory. Tellus,23, 404–425.

  • Lilly, D. H., 1966: On the instability of Ekman boundary layer flow. J. Atmos. Sci.,23, 481–494.

  • Louis, J. F., 1979: A parametric model of vertical eddy fluxes in the atmosphere. Bound.-Layer Meteor.,17, 187–202.

  • Mason, P. J., 1989: Large eddy simulation of the convective atmospheric boundary layer. J. Atmos. Sci.,46, 1492–1516.

  • Miura, Y., 1986: Aspect ratios of longitudinal rolls and convection cells observed during cold-air outbreaks. J. Atmos. Sci.,43, 26–39.

  • Müller, G., 1995: Mesoskalige Zellularkonvektion in Abhängigkeit von unterschiedlichen physikalischen Prozessen und synoptischen Randbedingungen: Numerische simulationen. Ph.D. dissertation, Berichte aus dem Zentrum für Meeres- und Klimaforschung, Hamburg, No. 16, 144 pp. [Available from Meteorologisches Institut, Bundesstr. 55, Hamburg D-20146, Germany.].

  • ——, and A. Chlond, 1996: Three-dimensional numerical study of cell broadening during cold-air outbreaks. Bound.-Layer Meteor.,81, 289–323.

  • Raasch, S., 1990: Numerical simulation of the development of the convective boundary layer during a cold-air outbreak. Bound.-Layer Meteor.,52, 349–375.

  • Rao, G.-S., and E. M. Agee, 1996: Large eddy simulation of turbulent flow in a marine convective boundary layer with snow. J. Atmos. Sci.,53, 86–100.

  • Rosmond, T. E., 1973: Mesoscale cellular convection. J. Atmos. Sci.,30, 1392–1409.

  • Tiedtke, M., 1989: A comprehensive mass flux scheme for cumulus parameterization in large-scale models. Mon. Wea. Rev.,117, 1779–1800.

  • Walter, B. A., 1986: The mesoscale organization, dynamics, and evolution of the marine planetary boundary layer during cold-air outbreaks. Ph.D. dissertation, University of Washington, 200 pp. [Available from University Microfilms, 300 North Zeeb Rd., Ann Arbor, MI 48106.].

  • Wayland, R. J., and S. Raman, 1989: Mean and turbulent structures of a baroclinic marine boundary layer during the 28 January 1986 cold-air outbreak (GALE 86). Bound.-Layer Meteor.,48, 227–254.

  • Fig. 1.

    National Oceanic and Atmospheric Administration infrared satellite imagery at 1249 UTC 25 March 1993, showing a cold-air outbreak off the ice edge near Spitsbergen. Linear cloud patterns change to cellular cloud structures with increasing distance from the ice edge. The flight pattern on this day is marked by white lines.

  • Fig. 2.

    Geographical location of the field experiments ARKTIS 1991 and ARKTIS 1993.

  • Fig. 3.

    Three-dimensional display of a typical flight pattern of the two research aircraft Falcon-20 and Dornier-128 during aircraft missions in cold-air outbreaks during the field experiments ARKTIS 1991 and ARKTIS 1993.

  • Fig. 4.

    Vertical profiles of potential temperature θ, specific humidity q, and wind components u and υ parallel and normal to mean wind direction (DD = 10°) measured at 1209 UTC 24 March 1993 at 78.4°N and 6.1°E at a distance of 160 km from the ice edge. The relevant basic flow parameters are indicated: boundary layer top h; cloud base hcb; air–sea differences Δθas, Δqas; vertical differences across the boundary layer Δθbl, Δqbl, Δubl, Δυbl and between h and h + 300 m, Δθct, Δqct, Δuct, Δυct, as well as of inflection point height hip in the u-wind profile.

  • Fig. 5.

    Time series of downwelling longwave radiation L↓, temperature T, water vapor mixing ratio m, and longitudinal (u), lateral (υ), and vertical (w) wind component measured during a crosswind flight leg at 92 m on 24 March 1993 within a field of roll convection. The centers of the overlying cloud streets are marked by thin vertical lines. Time marks are every 10 s corresponding to a distance of 1 km.

  • Fig. 6.

    Variance spectra SLL of longwave radiation weighted by frequency f and calculated from time series measured during four vertically staggered crosswind flight legs on 24 March 1993 within a field of roll convection. Frequencies f1 and f2 mark the limits of the roll-scale wavelength range. Frequency f can be converted to wavelength λ by λ = c/f with the aircraft speed of c = 100 m s−1.

  • Fig. 7.

    Variance spectra Saa (a = u, υ, w, θ, m) weighted by frequency f and calculated from the time series shown in Fig. 5. Frequencies f1 and f2 mark the limits of the roll-scale wavelength range and are taken from Fig. 6.

  • Fig. 8.

    (a) Wavelength λ vs height h of the secondary flow pattern and (b) aspect ratio λ/h vs ice edge distance Δx measured by aircraft in 20 cases during the field experiments ARKTIS 1991 and 1993. Encircled numbers mark the day of observation (see Table 1). Dashed lines connect cases observed at different Δx on the same day. Thin straight lines in (a) mark the λ/h bounds of roll patterns and cell patterns. The curved line in (b) represents the least squares logarithmic fit: λ/h = 5.5 + 1.7 lnΔx, where Δx is in units of 100 km.

  • Fig. 9.

    Sketch of the secondary flow circulation in the crosswind vertical y–z plane. Flight levels at the lowest permitted height (90 m) and in the middle of the boundary layer are indicated by dashed lines.

  • Fig. 10.

    Total and secondary flow–scale vertical fluxes of (a) sensible heat, wθ; (b) latent heat, mw; (c) longitudinal momentum, wu;and (d) lateral momentum, wυ. The fluxes are normalized as described in section 4c and are classified with respect to the convective pattern form in four categories: rolls (small), rolls (normal), rolls/cells, and cells. Numbers at the curves represent the day of observation (see Table 1). Thin horizontal lines mark cloud base and cloud top.

  • Fig. 11.

    Total (thick lines) and secondary flow–scale (dashed lines) vertical fluxes of latent heat, wm (above), and sensible heat, wθ (below), in regions of different convective patterns: rolls (small), rolls (normal), rolls/cells, and cells. The curves represent averages of the corresponding curves in Fig. 10. Thin horizontal lines mark cloud base and cloud top.

  • Fig. 12.

    Vertical profiles of longitudinal wind component u and lateral wind component υ relative to the corresponding values at 90-m height. Numbers indicate the day of observations (see Table 1). Horizontal lines represent cloud base and cloud top. The heights are scaled according to (4a), (4b).

  • Fig. 13.

    Vertical profiles of longitudinal and lateral wind components u and υ averaged for each convective category. Short horizontal lines represent standard deviation and long horizontal lines represent cloud base and cloud top. The heights are scaled according to (4a), (4b).

  • Fig. 14.

    Sketch of two superimposed cloud patterns observed during several flights within roll convection. The convective pattern along the x direction originates from thermal instability due to surface heating and is organized by wind shear and curvature of the u-wind component while the convective pattern along the y direction is supposed to originate from the inflection point instability due to an inflection point in the u-wind component close to the top of the boundary layer.

  • Fig. 15.

    Secondary flow–scale kinetic energy ½v2Rh integrated over the depth h of the boundary layer vs Rayleigh number Ra. The numbers mark the day of observation (see Table 1). The full line represents the least squares power law fit Ekin,R = 3.69Ra0.40. The dashed lines connect the two observations of increasing Ekin,R with increasing Ra on those five of the seven days when observations were made at two different distances Δx along a streamline from the ice edge.

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 1120 495 155
PDF Downloads 531 181 18