1. Introduction
Thermals—coherent, isolated, quasi-spherical regions of upward-moving buoyant fluid—are a common feature of convective atmospheric flows. A key characteristic of thermals is the rate at which they increase in size as they ascend owing to entrainment of the surrounding fluid. Assuming thermal shape is self-similar (meaning that thermals do not change shape over time), dimensional analysis shows that thermal radius R is proportional to thermal top height zt, that is, dR/dzt is constant1 (e.g., Scorer 1957). (Note that all symbols used in the paper are defined in the appendix.) Numerous laboratory and numerical modeling studies have supported this basic scaling (e.g., Scorer 1957; Richards 1961; Bond and Johari 2005; Zhao et al. 2013; Lai et al. 2015; Lecoanet and Jeevanjee 2019, hereinafter LJ2019; McKim et al. 2020; Morrison et al. 2021).
The rate of increase in R is closely related to the entrainment rate of thermals. From LJ2019, a thermal net fractional entrainment rate is defined as ϵ ≡ d(lnV)/dzt, where V is the thermal volume. Combined with self-similarity, this gives ϵ = 3α/R, where α ≡ dR/dzt. We emphasize that ϵ in this case is a net fractional entrainment rate because thermal volume is impacted by both entrainment (inflow of environmental fluid) and detrainment (outflow of thermal fluid). However, LJ2019 showed that detrainment is negligible for both laminar and turbulent dry, initially spherical thermals in a neutrally stable environment. Thus, ϵ provides a close approximation for total entrainment in such conditions. An entrainment efficiency can also be defined as e ≡ ϵR, which gives e = 3α for self-similar thermals.
Thermals entrain by a process of drawing in fluid mainly from below the thermal (e.g., LJ2019; Zhao et al. 2013; McKim et al. 2020; Morrison et al. 2021). As a thermal spins up, buoyancy becomes concentrated near the center of rotation in the thermal’s toroidal circulation (i.e., vortex ring core); see Fig. 1 for a schematic of thermal structure. As a result, there is baroclinic generation of buoyancy on the outside edge of the vortex ring and destruction on the inside edge that lead to a spreading of the vortex (McKim et al. 2020). Moreover, without buoyant fluid present along the thermal’s vertical axis, circulation is nearly constant. This implies a basic constraint on the spreading rate of thermals following the principle of momentum conservation (Turner 1957). Specifically, buoyant vortex rings (which form the core of thermals) must expand over time to conserve momentum, with the rate of spread determined by the thermal-integrated buoyant forcing and the circulation. McKim et al. (2020) combined the buoyant vortex ring argument of Turner (1957) with the thermal’s vertical momentum equation to derive an analytic model for the vertical velocity of thermal top wt, R, and buoyancy B at any time past spinup that does not rely on empirically determined parameters, provided wt, R, and B are known at the time when the thermal is spun up.
While the basic mechanism of thermal entrainment and spreading is well understood, factors controlling the spreading rate are not. Lai et al. (2015) combined a relation between circulation, impulse (related to time-integrated buoyant forcing), and thermal spreading rate with an empirical power-law relation between normalized circulation and initial thermal aspect ratio Ar to predict α from Ar. They showed that variations in Ar for spheroidal thermals from ∼0.5 to 2 lead to substantial variability in α, from about 0.1 to 0.3. These results are consistent with laboratory experiments reporting a similar range of α (e.g., Scorer 1957; Escudier and Maxworthy 1973; Bond and Johari 2005, 2010; Zhao et al. 2013). A consensus from laboratory and numerical modeling studies is that α ≈ 0.12–0.18 for initially spherical thermals in an unstratified environment (e.g., LJ2019; Bond and Johari 2010; Zhao et al. 2013; Lai et al. 2015). Values are ∼0.2–0.3 for initially oblate thermals with Ar < 1 and smaller for prolate thermals with Ar > 1, ∼0.1–0.15 (see Fig. 17 of Lai et al. 2015). There is little sensitivity of α to initial aspect ratio for Ar > 2 (Bond and Johari 2005). Modifying other aspects of thermal initial conditions can also produce variability in α, such as having an initial circulation (Escudier and Maxworthy 1973). Note that α may also depend on the Reynolds number Re of the flow, although LJ2019 showed with direct numerical simulation (DNS) that the basic mechanism of entrainment is the same for laminar and turbulent thermals (Re of 630 and 6300, respectively), and α was only ∼20% higher for turbulent thermals. Their results indicate that turbulence is not necessary for entrainment and that the primary mechanism for entrainment is organized inflow controlled by the thermal’s buoyancy distribution.
The above discussion raises two important science questions. First, why do initially spherical thermals in an unstratified environment (and initially motionless) have α ≈ 0.15? Why this particular value, and what are the physical mechanisms explaining it? Second, why does the spreading rate of thermals as they rise (α) increase as their initial aspect (Ar) is decreased? To our knowledge, all previous studies have relied in some way on empirical constraint to obtain parameters, from either laboratory experiments or numerical modeling, at least during the spinup phase which is crucial for predicting α. In this study, we derive an expression for α as a function of Ar that does not rely on such empirically determined parameters. The goal is to predict α from Ar from the basic equations to provide a theoretical underpinning for understanding factors controlling the thermal spreading rate. The predicted values of α are compared to those obtained from numerical simulations of thermals over a range of Ar.
In the theoretical part, we first derive an analytic expression for thermal ascent rate wt from the nondimensional thermal momentum budget equation. We then use this expression to derive an analytic relation between α and the thermal spinup height zc (defined as the thermal top height when wt reaches a maximum), valid over a range of initial thermal aspect ratios. We show that zc also corresponds to the time for parcels initially near the thermal bottom to ascend through the thermal core to near the thermal top. This determines the time for buoyancy to be removed from the central thermal core by entrainment of nonbuoyant environmental fluid, after which circulation is nearly constant. This time scale for thermal spinup depends linearly on Ar and is determined by the thermal’s internal flow structure which is well modeled by Hill’s analytic spherical vortex (Hill 1894) even for nonspherical thermals. We show that the predicted values of wt, α, and zc are consistent with numerical simulations of buoyant thermals over a range of Ar from 0.5 to 2.
The paper is organized as follows. Section 2 provides a theoretical description of the problem and derivation of equations for wt and α. Section 3 gives a description of the numerical model and experimental design. Section 4 presents results from the numerical simulations and comparison of these results with theory. A summary and conclusions are given in section 5.
2. Theoretical description
a. Thermal momentum budget
We assume that detrainment is negligible and that thermal expansion incorporates fluid with w = 0. Thus, net entrainment is related only to the change in thermal volume (following the Boussinesq approximation). This assumption is well justified based on the DNS of dry thermals from LJ2019; see Morrison et al. (2022) for further discussion. It follows that we can write the entrainment term as
We approximate the dynamic pressure part of thermal drag using the standard drag equation divided by V to give a volume-averaged dynamic pressure drag force:
Solutions for wt can be obtained from (11), provided values of α, e, γ, Cd, and Cυ are known. Past literature has suggested α ∼ 0.05–0.3 (Lai et al. 2015), e ≈ 3α (LJ2019), Cd ≈ 0 (Morrison et al. 2022), meaning that γ is not relevant, and Cυ ∼ 0.5–0.8 (Tarshish et al. 2018). Examples of solutions to the analytic wt Eq. (11) using e = 3α, Cd = 0, Cυ = 2/3, and α ranging from 0.05 to 0.3 are shown in Fig. 2 (solid lines). With these parameter values, (11) gives a family of solutions that all exhibit a sharp increase of wt with height initially corresponding to a “slippery” regime when upward buoyant forcing is primarily balanced by vertical acceleration, followed by a slower decrease of wt corresponding to a “sticky” regime when weak buoyant forcing is balanced mainly by entrainment.2 This shape of the wt profile with the two distinct regimes of thermal evolution was discussed previously via analysis of numerical solutions (e.g., Wang 1971; Tarshish et al. 2018). In subsequent sections, we will determine constraints on values of α while also briefly exploring how e and Cυ vary with the initial thermal aspect ratio.
We can understand scaling behaviors in the slippery and sticky regimes via asymptotic analysis and expansion of (11), similar to the asymptotic analysis of Escudier and Maxworthy (1973) applied to their equation set for wt. For the slippery regime when zt < 1, we can expand (11) using Taylor series about zt = 0 and retain the first-order term to give
In contrast, when zt ≫ 1 for the sticky regime, 1 + αzt ≈ αzt and the second term on the right-hand side of (11) is negligible compared to the first term (for 2e/α + γCd/2 > 2, which is satisfied for typical values of 2e/α ≈ 6 and γCd ≈ 0), implying a scaling of wt with
b. Impulse, circulation, and thermal spreading rate
If Γ is known when the thermal is spun up, we can derive an analytic expression for α by combining (15) and (17) with the relation between a thermal’s impulse and wt (Akhmetov 2009; McKim et al. 2020):
These complications motivate an alternative approach described below that relates α to the thermal top height when spinup is achieved, corresponding to the thermal top height when wt reaches its maximum. This is in a similar vein as relating α (or e) to ζ, FB, Γ, Cυ, and m, but with much simpler functional dependencies allowing for a clear understanding of the variation of α with Ar.
c. Relationship between thermal spinup height and spreading rate α
We define a critical thermal top height zc separating the “slippery” and “sticky” regimes when dwt/dzt = 0 and wt is a maximum. During spinup, dwt/dzt > 0 as dΓ/dt > 0 owing to the presence of buoyant fluid along the thermal’s vertical axis. After buoyancy is eroded along the thermal’s vertical axis from entrainment of environmental fluid, Γ is constant following (13), and thus, dwt/dzt < 0 following the
d. Relationship between zc and Ar
As argued in the previous subsection, zc corresponds to the thermal top height when buoyant fluid along the thermal’s vertical axis is replaced by entrained environmental fluid (meaning circulation is approximately constant thereafter). This erosion of buoyancy in the thermal core occurs as nonbuoyant parcels are entrained near the thermal bottom and move upward relative to the thermal as a whole. Thus, we expect the time scale for loss of buoyancy along the thermal’s vertical axis to be equal to the time for parcels entrained near the thermal bottom to travel upward through the thermal.
e. Predicting σ from Hill’s analytic spherical vortex
The thermal aspect ratio Ar is specified from the initial conditions, leaving σ as the only unknown parameter in (25) to obtain α. This parameter is closely related to the thermal internal flow structure, which controls the rate of parcel ascent in the thermal core relative to the thermal as a whole. Lai et al. (2015) noted similarity of the flow structure of thermals to Hill’s vortex, particular for Ar = 2. They found that the analytic Hill’s vortex solution deviated more from numerical thermal simulations for smaller Ar, but noted “it can still give a fair prediction of flow field” for Ar as low as 0.5. In agreement with Lai et al. (2015), in section 4, we show a close correspondence of vertical velocity profiles along the central axis in numerically simulated thermals to Hill’s vortex for initial Ar of 1 and 2, with more deviation but still fairly similar w profiles for Ar = 0.5.
A parcel initially at the bottom of Hill’s vortex will rise at the same rate as the vortex since u = 0 and w = W at this location (i.e., it is a stagnation point in the vortex-relative flow). However, a parcel initially just above the vortex bottom at r = 0 will rise relative to the vortex as a whole. Thermals, owing to their buoyancy, entrainment, and nonsteady behavior, do not have such stagnation points, and parcels initiated at the thermal bottom rise through the thermal depth as demonstrated by the simulations in section 4. Thermal flow is similar to Hill’s vortex in the interior. Thus, although parcel ascent differs between thermals and Hill’s vortex near the top and bottom boundaries, it is similar in the interior with an acceleration toward the center followed by a deceleration above.
Because of the stagnation points in Hill’s vortex, we cannot use it directly to estimate the Lagrangian time scale for parcel ascent starting from the thermal bottom. However, given similarity of the interior flow between thermals and Hill’s vortex, a rough approximation is to replace the Lagrangian time-mean w along the parcel’s path with the Eulerian vertical-mean w from Hill’s vortex:
Following discussion in Lai et al. (2015), the flow field of the Norbury vortex family (Norbury 1973), which generalizes Hill’s vortex to variable ring vortex thickness, may be closer to the thermal simulations with varying Ar. Similarly, the O’Brien (1961) analytic spheroidal vortex model might give a better description of the flow for spheroidal thermals. However, these models are steady state and also have stagnation points. Since Hill’s vortex provides a reasonable description of the interior thermal flow over a range of Ar, we use it to constrain σ following the discussion above.
3. Description of the numerical simulations
a. Model description and experimental design
We utilize the Cloud Model 1 (CM1) fluid flow model to numerically simulate thermals with varying initial Ar. CM1 is a nonhydrostatic model which has been widely used to simulate idealized atmospheric flows. Here, we use the incompressible Boussinesq configuration to solve the filtered Navier–Stokes equations similar to the large-eddy simulation (LES) configuration in Morrison et al. (2022). Prognostic variables are the 3D components of flow velocity and potential temperature perturbation θ′, although near-axisymmetry of the model fields is retained. Buoyancy B is obtained by gθ′/θ0, where θ0 is a constant background θ of the fluid environment. As noted by Morrison et al. (2022), in this framework, prognosing θ′ is equivalent to prognosing B itself. Simulations are nondimensionalized using a length scale equal to the radius of the initial thermal R0 (the radius of the initial buoyancy perturbation) and a time scale given by
The initial Ar of thermals is varied from 0.5 to 2, similar to the range from Lai et al. (2015). As we show in section 4, this produces a wide spread of α (∼0.08–0.25). Thermals are initiated by adding a buoyancy perturbation B0 uniformly within a spheroidal volume having a horizontal radius of R0 and a vertical radius of ArR0. To minimize the impacts of boundary conditions, the initial buoyancy perturbations are centered at a height of 4R0, and the horizontal domain width is ≥16R0 and the vertical domain height is 64ArR0 (64 times the initial vertical thermal radius). The model grid is isotropic in all three directions with a grid spacing ΔLm equal to 0.1ArR0. Since the initial Ar varies from 0.5 to 2, ΔLm ranges from 0.05R0 to 0.2R0. The time step is 0.0362 times the time scale
Configuration details for the CM1 simulations presented in this paper.
In this study, we use LES applied to the filtered Navier–Stokes equations instead of DNS to retain a close connection to atmospheric modeling, particularly modeling of dry and moist thermals in the planetary boundary layer and convective clouds in which DNS is not possible given the huge O(109) Reynolds numbers involved. The LES framework is also consistent with our previous work on dynamic drag of dry buoyant thermals (Morrison et al. 2022) and similar to previous thermal simulations of Lai et al. (2015). The subgrid-scale (SGS) mixing follows a Smagorinsky-type approach as implemented by Stevens et al. (1999, see their appendix B, section b). The SGS mixing length is set to ΔLm. Because the dissipation scale (the model’s filter scale) is a relatively large fraction of the thermals’ radii, the resolved scale flow is smooth and thus appears laminar. The resulting thermal evolution and internal flow structure of the simulated thermals is remarkably similar to the DNS of initially spherical laminar thermals in LJ2019 (Re = 630). Results across the range of Ar are close to those of Lai et al. (2015), who also numerically solved the filtered (discretized) Navier–Stokes equations but using a k–δ turbulence closure (Launder and Spaulding 1974), where k is the resolved kinetic energy and δ is the energy dissipation rate. Our simulations are integrated forward in time until the thermal top (as defined in section 3b) reaches a height of 15R0 above the initial thermal top (i.e., top of the initial buoyancy perturbation). To investigate the internal thermal flow characteristics, particularly the time for ascent of a parcel through the thermal, each simulation includes forward trajectories for a parcel placed at the thermal bottom at the initial time. We use the built-in parcel trajectory calculations in CM1 which are done during the model integration using linear interpolation of the flow field at each model time step.
b. Analysis methodology
Thermal boundaries must first be identified and tracked in order to analyze thermal behavior including spreading rate. We use a method similar to LJ2019 and Morrison et al. (2022). At each output time (at an interval of 0.542 times the time scale
Other quantities of interest are 1) vorticity, which is calculated directly from the velocity field using centered finite differencing, and 2) buoyant and dynamic components of perturbation pressure, output directly from the model as described in Morrison et al. (2022).
4. Analysis of numerical simulations
Overall structure and evolution is similar for all of the simulated thermals. Starting from rest, rapid spinup ensues owing to vorticity generation from the thermals’ buoyancy distributions. The thermals spread outward as they rise and entrain the surrounding fluid. Spinup of the thermals (after which circulation is nearly constant) occurs when a parcel initially placed at the thermal bottom rises to near proximity of the thermal top. Here, we calculate the critical height zc as the thermal top height when spinup is achieved, rather than directly from the height where dwt/dzt = 0 and wt is maximum because dwt/dzt is rather noisy. Nonetheless, zc calculated from the parcel trajectories matches well with broad maxima in wt as shown later.
In accordance with the theory presented in section 2, zc ranges from about 1 to 6 as Ar is varied from 0.5 to 2 (Table 2). After spinup, when zt > zc in the “sticky” regime, the thermals continue to expand by entraining environmental fluid, but their overall flow structure is fairly steady. The thermals undergo a slow deceleration (relative to the faster acceleration during spinup) with wt roughly proportional to t−1/2 (and thus also proportional to
Time-averaged α ≡ dR/dzt, b ≡ e/α, virtual mass parameter Cυ, ratio of parcel to thermal-time-averaged vertical velocity σ, circulation Γ, and critical spinup height zc from the simulations with varying Ar. Note that b is obtained from the ratio of time-averaged e to time-averaged α. Because of some noise in calculating thermal velocity directly, σ is derived from (23) using zc obtained from the simulations as described in the text. α, e, and Cυ are calculated as time averages over the full simulation period, whereas Γ values are time-averaged after spinup to the end of the simulations.
Figure 4 shows vertical cross sections through the thermal center of B, w, horizontal vorticity in the y direction ηy, and streamfunction ψ after spinup, when thermal top height is at approximately zc + 2R0. Thermal flow features well documented by previous studies are seen in the figure. These include toroidal circulations with rotation centers near the thermal vertical midpoint, buoyancy concentrated near these rotation centers, and downward motion (in an absolute sense and relative to wt) along the thermal periphery. Although buoyancy is almost entirely swept away from the thermal core (along the vertical axis at X = 0) for the Ar = 0.5 and 1 simulations, some positive buoyancy remains in the core when Ar = 2. There is also fluid with B > 0 and ηy ≠ 0 below the thermal in this simulation (Figs. 4e,f). This occurs because not all of the initially buoyant fluid is taken into the thermal’s vortex ring (toroidal circulation) when the aspect ratio is large, a result also noted by Lai et al. (2015). This behavior can be described by the “formation number” (Gharib et al. 1998), which is related to the maximum vorticity that can be incorporated into a vortex ring before it “pinches off” from a trailing stem. Earlier work showed a formation number of 4–5 for vortex rings (Gharib et al. 1998; Wang et al. 2009), whereas Lai et al. (2015) found a somewhat lower formation number of ∼2, consistent with our results.3 Despite the presence of a trailing stem of weakly buoyant fluid in the Ar = 2 simulation, buoyancy in the core is small relative to that near the rotation centers, and as detailed later, the theoretical relations between zc, α, and Ar proposed in section 2 still well describe behavior of this simulation. We suspect that further increases in Ar would lead to greater deviation with the theory. Indeed, Lai et al. (2015) showed little change in α as Ar was increased beyond 2, likely because of the inability of such thermals to incorporate all of the initially buoyant fluid. The behavior of these thermals instead resembled a starting plume, consistent with the numerical results of Bond and Johari (2010).
Differences in thermal aspect ratio with varying initial Ar persist beyond spinup, although these differences are reduced compared to the initial Ar. The thermals with initial Ar ≥ 1 become more flattened (smaller aspect ratio) during spinup. At the times shown in Fig. 4, the Ar = 2 simulation has an aspect ratio just slightly larger than 1, while that for Ar = 1 is about 0.75 and that for Ar = 0.5 is about 0.6. Different thermal aspect ratios among the simulations are reflected by variability in time-averaged values of Cυ (virtual mass parameter, see section 2); see Table 2. Here, Cυ is calculated at each model output time directly from the buoyancy and buoyant pressure forcing averaged over the thermal volume. Larger initial aspect ratios are associated with larger Cυ, consistent with results from Tarshish et al. (2018). There is also an overall decrease in Cυ over time during spinup as the thermals flatten, particularly for the simulations with Ar > 1. Changes in thermal shape during spinup also lead to deviation in b from the value for self-similar thermals (b = 3). The Ar = 2 thermal has the largest deviation, with b ≈ 2.06, which is consistent with it experiencing the greatest change in aspect ratio during spinup, whereas b ranges from ∼2.4 to 2.7 for the other simulations.
Thermal behavior during spinup is illustrated in Fig. 5, which shows vertical cross sections of B, w, ηy, and ψ in the same format as Fig. 4 except during the spinup period for the Ar = 1 simulation. Cross sections are shown in nondimensional time increments of 1.1 between t = 1.6 and 4.9. For context, the thermal top reaches zc at t ≈ 3.7. The basic mechanism of spinup is similar for all the runs. Consistent with the discussion in section 2d, entrainment occurs as environmental fluid is swept into the thermal from below in the convergent flow. This appears as a “bite” taken from the buoyancy field from below and occurs because thermal-relative vertical velocities are strongest in the thermal core. Baroclinic vorticity generation is concentrated along the edge of the buoyancy field where there are large horizontal buoyancy gradients. Once the buoyancy field is deformed and starts to wrap around the vortex core (i.e., the center of rotation), baroclinic generation and destruction of vorticity drives a spreading of the thermal in the manner outlined by McKim et al. (2020) and Morrison et al. (2021). Flattening of the thermal during spinup is also evident in Fig. 5.
In all simulations, the thermals’ internal flow structures consist of thermal-relative ascent in the core, with strongest ascent along the vertical axis and descent along the periphery. This flow pattern strongly resembles Hill’s analytic spherical vortex. To illustrate this point further, Fig. 6 compares w profiles along the thermals’ vertical axis from the simulations with Ar of 0.5, 1, and 2 with w profiles at the vertical axis from Hill’s vortex given by (26) and (27). This is similar to the comparison of w profiles from thermal simulations with Hill’s vortex in Lai et al. (2015, Fig. 12 therein). Simulation results here are shown at the time of spinup when the thermal top is at zc. Profiles from the simulations are normalized by the maximum w with height normalized by the thermal depth; thermal bottom and top heights are set to −1 and 1, respectively. Correspondingly, a = 1 in (26) and (27) for the Hill’s vortex w profile. All of the simulations produce similar w profiles as Hill’s vortex, with the Ar = 1 being closest. There is also a close correspondence of the Ar = 2 simulation with Hill’s vortex, with greater deviation for Ar = 0.5. Overall, these results support the discussion in section 2e on the validity of approximating σ for thermals from Hill’s vortex.
Differences in spreading rate α ≡ dR/dz among the simulated thermals are seen in Fig. 7a, which shows thermal radius R as a function of zt for the simulations with Ar of 0.5, 1, and 2. The increase of R with zt is clearly greater as Ar is decreased, with α about 3 times larger in the Ar = 0.5 simulation compared to Ar = 2. Although R is somewhat noisy, the overall spreading rates are nearly constant with zt (seen by the dotted lines) consistent with similarity theory.
A comparison of simulated wt as a function of zt with solutions to the analytic wt equation, Eq. (11), is shown in Fig. 7b. The analytic wt are obtained using mean values of Cυ, b, and α from each simulation (Table 2). The overall behavior of wt is similar among the simulations, with a sharp increase during spinup followed by a slower decrease after spinup. The analytic wt are close to the simulated values for each simulation (compared the dotted and solid lines in Fig. 7). Larger values of α in the Ar = 0.5 simulation correspond to faster spinup and lower height of maximum wt (critical height zc) compared to the Ar = 1 and especially Ar = 2 simulations with smaller α. Thus, zc increases with Ar consistent with the theory in section 2d.
Thermal behavior during spinup for the simulations with Ar of 0.5, 1, and 2 is further illustrated in Fig. 8, which shows time series of thermal top height zt and vertical velocity wt, circulation Γ, and vertically integrated core buoyancy (
Values of zc from the simulations, estimated from zp and zt as described above, are compared to the theoretical linear zc–Ar relation (23) using σ = 2 from Hill’s analytic vortex (see section 2e) and using the average σ = 1.80 from the simulations (Table 2) in Fig. 9a. The simulated and theoretical zc values are similar, although the Ar = 2 simulation deviates more substantially from the theoretical relations. Reasons for this deviation are unclear but might be explained by the buoyant fluid entrained into the thermal from below in this simulation, leading to some buoyancy remaining along the thermal’s vertical axis even after spinup. The theoretical σ ≈ 2 derived from the Eulerian mean w of Hill’s vortex (see section 2e) is fairly close to σ values obtained directly from the simulations (within 10% except for the Ar = 2 simulation), though somewhat larger. The simulated values range from 1.80 to 1.95 for Ar < 2 but are slightly smaller (≈1.63) for Ar = 2 (see Table 2).
A direct comparison of the simulated and theoretical values of spreading rate α is shown in Fig. 9b. Theoretical values are obtained from 1) Eq. (19) using zc and b derived from the simulations (Table 2), 2) Eq. (20) using zc derived from the simulations and b = 3 following self-similarity, 3) Eq. (25) using the average σ = 1.80 from the simulations to predict zc, and 4) Eq. (31) which calculates α from zc predicted using σ = 2 from Hill’s analytic vortex. All of the theoretical calculations for α give similar results as the simulations. The simulations show a sharp decrease of α with increasing Ar that follows an approximate
5. Discussion
Overall, the simulations and theory are in reasonable agreement regarding thermal top height at spinup zc and thermal spreading rate α and how they vary with initial aspect ratio Ar. Our results indicate a nearly linear relation between zc and Ar (though with greater deviation for the Ar = 2 simulation) and an inverse relation between α and Ar (
This explains why larger α is associated with smaller Ar but does not by itself explain the physical mechanism. A key question, therefore, is what is the mechanism driving the increase in α as Ar is reduced? With small Ar,
This work also provides a concise explanation for why initially spherical thermals (Ar = 1) have α ≈ 0.15 (for an unstratified, neutrally stable environment). This value of α is intrinsically linked to the time scale for sweeping out of the buoyancy along the thermal’s vertical axis and hence thermal spinup, which itself depends on the ratio (σ) of time-averaged wp to wt. The thermals’ internal flow structures are similar to Hill’s analytic spherical vortex, implying σ ≈ 2 and in turn constraining the proportionality constant in the
An interesting feature is that, in a given simulation, α is similar before and after thermal spinup. This is evident directly from the simulations (profiles of R in Fig. 7a, although they are somewhat noisy, and the vertical cross sections of thermal properties during spinup in Fig. 5), as well as indirectly by closeness of the simulated and theoretical w profiles (Fig. 7b), the latter calculated assuming constant α. Thus, α values are “locked in” early in the simulations, and they depend strongly on the initial conditions. Why is α similar during spinup and after? A possible explanation is that circulation is small early in the simulations, while at the same time, entrainment has only just begun to erode buoyancy in the thermal core. This means that buoyancy gradients and hence baroclinic generation and destruction of vorticity near the central core are weak (vorticity generation being concentrated more along the thermal boundary). However, because circulation and vorticity near the vortex core are also weak, the net result is a similar thermal spreading rate compared to later when both baroclinic generation/destruction of vorticity and circulation are stronger. Moreover, α is the change in R with zt, and small w during early spinup means that a small spreading rate in time is associated with a relatively larger α.
We also note that α is somewhat larger for turbulent compared to laminar thermals; LJ2019 and Morrison et al. (2022) found ∼20% and 40% larger values for turbulent thermals, respectively. At high Reynolds number, turbulent stresses lead to a spindown of circulation after thermal spinup such that dΓ/dt < 0 (Nikulin 2014; McKim et al. 2020). All else equal, dΓ/dt < 0 implies a larger spreading rate following the impulse–circulation relation (16). Nikulin (2014) developed an analytic expression for α as a function of Γ, FB, an empirical parameter β (encapsulating ζ, Cυ, and m), and an empirical proportionality constant characterizing the impact of turbulent stresses. Using parameter values deduced from experimental data, they suggested a ∼3% increase in α from turbulent stresses. However, this study did not consider the effects of turbulent stresses on thermals starting from rest. Reduced circulation from turbulent stresses during spinup might explain the order-of-magnitude larger impact on α found by LJ2019 and Morrison et al. (2022), as both studies simulated thermals that were initially motionless. The hypothesis that turbulent stresses during spinup cause most of the differences in α between laminar and turbulent thermals is consistent with Fig. 3 in McKim et al. (2020), which shows that the turbulent case has ∼30% smaller Γ at the time of spinup relative to the laminar case.
6. Summary and conclusions
This study investigated the spreading rate α and entrainment behavior of dry, buoyant thermals with varying initial aspect ratio Ar. An expression was derived for the nondimensional thermal ascent rate wt as a function of thermal top height zt from the thermal w momentum budget. From this expression, we defined a critical thermal top height zc where dwt/dzt = 0. The height zc corresponds to the thermal top height when buoyancy is eroded along the thermal’s vertical axis from entrainment of nonbuoyant environmental fluid (with thermal circulation approximately constant thereafter). We then analytically solved dwt/dzt = 0 to derive an expression relating α and zc. In turn, zc depends on Ar and the ratio σ of the mean vertical velocity of a parcel rising from thermal bottom to near its top along its vertical axis to wt. By approximating the thermal flow similarly to Hill’s analytic spherical vortex, it was estimated σ ≈ 2. In this way, we derived an analytic expression for α that depends inversely on Ar.
Numerical simulations of thermals with Ar varying from 0.5 to 2 were analyzed and compared to the theoretical expressions. The analytic formulation for wt well matched the thermal simulations over the range of Ar. Values of α calculated directly from the simulations were also close to the theoretical α over the range of Ar. Consistent with the theory, increasing Ar led to slower spinup owing to an increase in distance (relative to the thermal radius) for parcels to travel from thermal bottom to near top, meaning that core buoyancy was eroded more slowly by entrainment. Values of σ were similar among the simulations and ranged from 1.63 to 1.95, somewhat less than the theoretical σ ≈ 2 based on the flow similarity between the thermals and Hill’s vortex. This work also provided an explanation for why initially spherical thermals (Ar = 1) have α ≈ 0.15, which occurs because of the similarity of thermal flow to Hill’s vortex. This gives σ ≈ 2 and constrains the proportionality constant in the
This study has elucidated factors controlling the spreading rate of dry buoyant thermals. This work is relevant to buoyantly driven atmospheric flows, especially those with a localized pulse source of buoyancy or steady source that leads to a chain of multiple thermals. In particular, numerous studies have noted the importance of buoyant thermals for cumulus convection in the atmosphere (e.g., Blyth et al. 2005; Damiani et al. 2006; Sherwood et al. 2013; Romps and Charn 2015; Hernandez-Deckers and Sherwood 2018; Morrison et al. 2020; Peters et al. 2020). Spreading rates of dry thermals may also indirectly impact cumulus entrainment rates by influencing the size of thermals at cloud base (Mulholland et al. 2021). Although Vybhav and Ravichandran (2022) suggested similar growth rates for dry and moist (cloud) thermals, Morrison et al. (2021) found that the spreading rate of moist thermals was almost a factor of 2 smaller than dry thermals for conditions typical of cumulus convection in the lower and middle troposphere. It is unclear how results from the current study might translate to cumulus thermals, given the impact of latent heating and cooling on their buoyancy distributions. Moreover, for buoyantly driven atmospheric flows at scales of interest, dry and moist thermals are generally turbulent. Nikulin (2014) suggested that the effects of turbulent stresses can be considered as an additional term leading to a small increase in α. This is supported by the recent numerical modeling studies of LJ2019 and Morrison et al. (2022), although they demonstrated an order-of-magnitude larger impact on α than Nikulin (2014) (∼20%–40% versus a few percent). Future work should refine understanding of the entrainment behavior and spreading rates for dry and moist turbulent thermals.
Note that constant dR/dzt following self-similarity and dimensional analysis is valid when there are no other physical length scales. It follows that this scaling applies to dry thermals in an unstratified environment within an infinite domain.
Note that the sticky regime is primarily a balance between entrainment and buoyant forcing for dry thermals, whereas Romps and Charn (2015) identified a sticky regime for cloud thermals consisting mainly of a balance primarily between buoyant forcing and downward pressure gradient forcing.
Note that this difference may be explained in part because Gharib et al. (1998) defined formation number by the maximum vorticity incorporated into the vortex ring, while Lai et al. (2015) defined it by the maximum volume of fluid incorporated.
Acknowledgments.
HM was supported by the U.S. Department of Energy Atmospheric System Research Grant DE-SC0020104. The National Center for Atmospheric Research is sponsored by the National Science Foundation. We acknowledge high-performance computing support from Cheyenne (DOI:10.5065/D6RX99HX) provided by NCAR’s Computational and Information Systems Laboratory. We thank Dr. Steven Sherwood for helpful discussions and Dr. George Bryan for developing and maintaining CM1.
Data availability statement.
This study used CM1 version 20.1 (cm1r20.1) released on 25 August 2020. CM1 code and detailed documentation are available at https://www2.mmm.ucar.edu/people/bryan/cm1/. Configuration and namelist files for the CM1 simulations as well as analysis code can be made available upon request to the first author.
APPENDIX
List of Symbols
a |
Radius of Hill’s vortex |
A |
Thermal cross-sectional area |
Ar |
Initial thermal aspect ratio |
b |
Parameter defined by the ratio e to α |
B |
Buoyancy |
Bc |
Core buoyancy along the thermal’s vertical axis |
Beff |
Effective buoyancy |
Cd |
Dynamic drag coefficient |
Cυ |
Virtual mass parameter |
D |
Distance traveled by the vortex as a whole over time period Δt |
D0 |
Initial thermal vertical length |
e |
Entrainment efficiency |
E |
Momentum entrainment |
Eu |
Euler number |
f |
Fractional distance from vortex center where the parcel is initiated relative to radius a |
Fd |
Thermal-averaged pressure drag force |
Fr |
Froude number |
FpB |
Thermal-averaged buoyant pressure drag force |
FpD |
Thermal-averaged dynamic pressure drag force |
g |
Gravitational acceleration |
I |
Fluid impulse |
Iz |
Fluid impulse in the z direction |
Unit vector in the vertical | |
k1 |
Integration constant |
m |
Shape parameter defined as the ratio of V to R3 |
p |
Pressure |
Unit vector normal to the thermal’s surface | |
u |
Fluid velocity vector |
uaxi |
Regridded radial velocity in cylindrical coordinates |
ub |
Displacement rate of thermal boundary |
ue |
Effective entrainment velocity |
r |
Radial direction in axisymmetric coordinates |
R |
Thermal radius |
Re |
Reynolds number |
Rυ |
Ring vortex radius |
S |
Region defined by circuit passing through the thermal core and returning through the ambient fluid |
t |
Time |
V |
Thermal volume |
w |
Fluid vertical velocity |
W |
Velocity of Hill’s vortex |
waxi |
Regridded vertical velocity in cylindrical coordinates |
wp |
Vertical velocity of a parcel along its Lagrangian path |
wt |
Vertical velocity of thermal top |
z |
Height |
zbs |
Height at the bottom of region S |
zc |
Thermal top height at spinup |
zt |
Height of thermal top |
zts |
Height at the top of region S |
α |
Rate of increase in thermal radius with height as the thermal rises, equivalent to dR/dz |
γ |
Thermal shape parameter defined as the ratio AR/V |
Γ |
Thermal circulation |
ΔLm |
Grid spacing of the numerical model |
Δt |
Time for parcel to travel from near vortex bottom to near its top |
ϵ |
Fractional entrainment rate |
ηy |
Horizontal vorticity in the y direction |
ωϕ |
Azimuthal vorticity |
Ω |
Region of space occupied by thermal |
ψ |
Streamfunction |
θ |
Potential temperature |
ρ |
Fluid density |
ρ0 |
Constant background fluid density |
σ |
Ratio of time-averaged vertical velocities of the parcel and thermal top |
τc |
Time scale for thermal top to reach zc |
ζ |
Ratio of ring vortex radius to thermal radius |
REFERENCES
Akhmetov, D. G., 2009: Vortex Rings. Springer, 151 pp., https://doi.org/10.1007/978-3-642-05016-9.
Batchelor, G. K., 2000: An Introduction to Fluid Dynamics. Cambridge University Press, 615 pp.
Blyth, A. M., S. G. Lasher-Trapp, and W. A. Cooper, 2005: A study of thermals in cumulus clouds. Quart. J. Roy. Meteor. Soc., 131, 1171–1190, https://doi.org/10.1256/qj.03.180.
Bond, D., and H. Johari, 2005: Effects of initial geometry on the development of thermals. Exp. Fluids, 39, 591–601, https://doi.org/10.1007/s00348-005-0997-1.
Bond, D., and H. Johari, 2010: Impact of buoyancy on vortex ring development in the near field. Exp. Fluids, 48, 737–745, https://doi.org/10.1007/s00348-009-0761-z.
Damiani, R., G. Vali, and S. Haimov, 2006: The structure of thermals in cumulus from airborne dual-Doppler radar observations. J. Atmos. Sci., 63, 1432–1450, https://doi.org/10.1175/JAS3701.1.
Escudier, M. P., and T. Maxworthy, 1973: On the motion of turbulent thermals. J. Fluid Mech., 61, 541–552, https://doi.org/10.1017/S0022112073000856.
Gharib, M., E. Rambod, and K. Shariff, 1998: A universal time scale for vortex ring formation. J. Fluid Mech., 360, 121–140, https://doi.org/10.1017/S0022112097008410.
Hernandez-Deckers, D., and S. C. Sherwood, 2018: On the role of entrainment in the fate of cumulus thermals. J. Atmos. Sci., 75, 3911–3924, https://doi.org/10.1175/JAS-D-18-0077.1.
Hill, M. J. M., 1894: On a spherical vortex. Philos. Trans. Roy. Soc., A185, 213–245.
Lai, A. C. H., B. Zhao, A. W.-K. Law, and E. E. Adams, 2015: A numerical and analytical study of the effect of aspect ratio on the behavior of a round thermal. Environ. Fluid Mech., 15, 85–108, https://doi.org/10.1007/s10652-014-9362-3
Launder, B. E., and D. B. Spaulding, 1974: The numerical computation of turbulent flows. Comput. Methods Appl. Mech. Eng., 3, 269–289, https://doi.org/10.1016/0045-7825(74)90029-2.
Lecoanet, D., and N. Jeevanjee, 2019: Entrainment in resolved, dry thermals. J. Atmos. Sci., 76, 3785–3801, https://doi.org/10.1175/JAS-D-18-0320.1.
McKim, B., N. Jeevanjee, and D. Lecoanet, 2020: Buoyancy-driven entrainment in dry thermals. Quart. J. Roy. Meteor. Soc., 146, 415–425, https://doi.org/10.1002/qj.3683.
Morrison, H., J. M. Peters, A. C. Varble, W. M. Hannah, and S. E. Giangrande, 2020: Thermal chains and entrainment in cumulus updrafts. Part I: Theoretical description. J. Atmos. Sci., 77, 3637–3660, https://doi.org/10.1175/JAS-D-19-0243.1.
Morrison, H., J. M. Peters, and S. C. Sherwood, 2021: Comparing growth rates of simulated moist and dry convective thermals. J. Atmos. Sci., 78, 797–816, https://doi.org/10.1175/JAS-D-20-0166.1.
Morrison, H., N. Jeevanjee, and J.-I. Yano, 2022: Dynamic pressure drag on rising buoyant thermals in a neutrally stable environment. J. Atmos. Sci., 79, 3045–3063, https://doi.org/10.1175/JAS-D-21-0274.1.
Mulholland, J. P., J. M. Peters, and H. Morrison, 2021: How does LCL height influence deep convective updraft width? Geophys. Res. Lett., 48, e2021GL093316, https://doi.org/10.1029/2021GL093316.
Nikulin, V. V., 2014: Analytical model of motion of turbulent vortex rings in an incompressible fluid. J. Appl. Mech. Tech. Phys., 55, 558–564, https://doi.org/10.1134/S0021894414040026.
Norbury, J., 1973: A family of steady vortex rings. J. Fluid Mech., 57, 417–431, https://doi.org/10.1017/S0022112073001266.
O’Brien, V., 1961: Steady spheroidal vortices—More exact solutions to the Navier-Stokes equation. Quart. Appl. Math., 19, 163–168, https://doi.org/10.1090/qam/137415.
Peters, J. M., H. Morrison, A. C. Varble, W. M. Hannah, and S. E. Giangrande, 2020: Thermal chains and entrainment in cumulus updrafts. Part II: Analysis of idealized simulations. J. Atmos. Sci., 77, 3661–3681, https://doi.org/10.1175/JAS-D-19-0244.1.
Richards, J. M., 1961: Experiments on the penetration of an interface by buoyant thermals. J. Fluid Mech., 11, 369–384, https://doi.org/10.1017/S0022112061000585.
Romps, D. M., and A. B. Charn, 2015: Sticky thermals: Evidence for a dominant balance between buoyancy and drag in cloud updrafts. J. Atmos. Sci., 72, 2890–2901, https://doi.org/10.1175/JAS-D-15-0042.1.
Scorer, R. S., 1957: Experiments on convection of isolated masses of buoyant fluid. J. Fluid Mech., 2, 583–594, https://doi.org/10.1017/S0022112057000397.
Sherwood, S. C., D. Hernández-Deckers, M. Colin, and F. Robinson, 2013: Slippery thermals and the cumulus entrainment paradox. J. Atmos. Sci., 70, 2426–2442, https://doi.org/10.1175/JAS-D-12-0220.1.
Stevens, B., C. H. Moeng, and P. P. Sullivan, 1999: Large-eddy simulations of radiatively driven convection: Sensitivities to the representation of small scales. J. Atmos. Sci., 56, 3963–3984, https://doi.org/10.1175/1520-0469(1999)056<3963:LESORD>2.0.CO;2.
Tarshish, N., N. Jeevanjee, and D. Lecoanet, 2018: Buoyant motion of a turbulent thermal. J. Atmos. Sci., 75, 3233–3244, https://doi.org/10.1175/JAS-D-17-0371.1.
Turner, J. S., 1957: Buoyant vortex rings. Proc. Roy. Soc. London, 239A, 61–75, https://doi.org/10.1098/rspa.1957.0022.
Vybhav, G. R., and S. Ravichandran, 2022: Entrainment in dry and moist thermals. Phys. Rev. Fluids, 7, 050501, https://doi.org/10.1103/PhysRevFluids.7.050501.
Wang, C. P., 1971: Motion of an isolated buoyant thermal. Phys. Fluids, 14, 1643–1647, https://doi.org/10.1063/1.1693659.
Wang, R. Q., A. W. K. Law, E. E. Adams, and O. B. Fringer, 2009: Buoyant formation number of a starting buoyant jet. Phys. Fluids, 21, 125104, https://doi.org/10.1063/1.3275849.
Zhao, B., A. W. Law, A. C. Lai, and E. E. Adams, 2013: On the internal vorticity and density structures of miscible thermals. J. Fluid Mech., 722, R5, https://doi.org/10.1017/jfm.2013.158.