Generation of Infrasound by Evaporating Hydrometeors in a Cloud Model

David A. Schecter NorthWest Research Associates, Redmond, Washington

Search for other papers by David A. Schecter in
Current site
Google Scholar
PubMed
Close
and
Melville E. Nicholls Cooperative Institute for Research in Environmental Sciences, University of Colorado, Boulder, Colorado

Search for other papers by Melville E. Nicholls in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

The dynamical core of the Regional Atmospheric Modeling System has been tailored to simulate the infrasound of vortex motions and diabatic cloud processes in a convective storm. Earlier studies have shown that the customized model (c-RAMS) adequately simulates the infrasonic emissions of generic vortex oscillations. This paper provides evidence that c-RAMS accurately simulates the infrasound associated with parameterized phase transitions of cloud moisture. Specifically, analytical expressions are derived for the infrasonic emissions of evaporating water droplets in dry and humid environments. The dry analysis considers two single-moment parameterizations of the microphysics, which have distinguishable acoustic signatures. In general, the analytical results agree with the numerical output of the model. An appendix briefly demonstrates the ability of c-RAMS to accurately simulate the infrasound of the entropy and mass sources generated by an equilibrating cloud of icy hydrometeors.

Corresponding author address: David A. Schecter, NorthWest Research Associates, 4118 148th Ave. NE, Redmond, WA 98052. Email: schecter@nwra.com

Abstract

The dynamical core of the Regional Atmospheric Modeling System has been tailored to simulate the infrasound of vortex motions and diabatic cloud processes in a convective storm. Earlier studies have shown that the customized model (c-RAMS) adequately simulates the infrasonic emissions of generic vortex oscillations. This paper provides evidence that c-RAMS accurately simulates the infrasound associated with parameterized phase transitions of cloud moisture. Specifically, analytical expressions are derived for the infrasonic emissions of evaporating water droplets in dry and humid environments. The dry analysis considers two single-moment parameterizations of the microphysics, which have distinguishable acoustic signatures. In general, the analytical results agree with the numerical output of the model. An appendix briefly demonstrates the ability of c-RAMS to accurately simulate the infrasound of the entropy and mass sources generated by an equilibrating cloud of icy hydrometeors.

Corresponding author address: David A. Schecter, NorthWest Research Associates, 4118 148th Ave. NE, Redmond, WA 98052. Email: schecter@nwra.com

1. Introduction

Recent observations and longstanding theoretical considerations suggest that a developing tornado has a detectable signature in the infrasound1 of a severe storm (Bedard 2005; Bedard et al. 2004; Georges 1976; Passner and Noble 2006; Szoke et al. 2004). To reliably distinguish the vortex signal from extraneous noise, we must improve our current understanding of the various mechanisms that produce infrasound in atmospheric convection. Without detailed observations of the acoustic sources within a storm, numerical modeling may be the best method of investigation.

On the other hand, standard cloud models were not designed for the study of atmospheric infrasound. Their capabilities and limitations in this area of research are not fully understood. Here, we consider a special version of the Regional Atmospheric Modeling System (RAMS) that has been customized to simulate acoustics (Nicholls and Pielke 2000). Cotton et al. 2003 provide an overview of standard RAMS, whereas details of the microphysics parameterizations can be found in a number of additional papers (Walko et al. 1995, 2000; Meyers et al. 1997; Saleeby and Cotton 2004). Appendix A briefly describes the customized version of the model, which is called c-RAMS hereinafter.

The literature contains some evidence that c-RAMS has an adequate foundation for modeling the infrasound of convective storms. Schecter et al. (2008) recently showed that c-RAMS can simulate the adiabatic generation of infrasound by the Rossby-like waves of substorm-scale vortices. Nicholls and Pielke (1994a,b, 2000) previously showed that c-RAMS can simulate the creation of low-frequency compression waves (30-min Lamb waves) by storm-scale heating. However, no prior study has verified that c-RAMS accurately simulates the infrasound that is generated by phase transitions of moisture in the 0.1–10-Hz frequency band. This critical frequency band is where severe storms are observed to produce abnormally strong and distinct signals (Bedard 2005). In theory, adiabatic vortex motions could account for the observations (Bedard 2005; Schecter et al. 2008). However, entropy and mass fluctuations associated with phase transitions may also contribute. A recent analytical study concluded that such fluctuations in the moist turbulence of a severe storm are likely to dominate other sources of 0.1-Hz infrasound (Akhalkatsi and Gogoberidze 2009).

In this paper, we illustrate the fundamental mechanism by which phase transitions produce infrasound in c-RAMS. We focus on a conceptually simple paradigm—the evaporation of an isolated, homogeneous cloud of water droplets. Analytical expressions are derived for the acoustic emissions in dry and humid environments. Under sufficiently dry conditions, the evaporation may occur in a few seconds or less. Successful comparison of the analytical results to numerical experiments verifies that the practical output of c-RAMS agrees with the theoretical thermo–acoustics of the model. Such verification helps justify using c-RAMS for future numerical studies of infrasound generated by convective storms.

The remainder of this paper is organized as follows: section 2 presents a relatively simple theory for the acoustic emissions of an evaporating cloud in a dry environment, using two different single-moment microphysics parameterizations. Section 3 compares the analytical results of section 2 with numerical simulations. Section 4 presents a theory for the acoustic emissions of an evaporating cloud in a humid environment, and compares the results with numerical simulations. Section 5 contains a summary and concluding remarks. The appendixes discuss our customization of RAMS, the infrasound of icy hydrometeors, and theoretical subtleties regarding evaporation under humid conditions.

2. Theory of infrasound generated by evaporation in a dry environment

This section derives analytical formulas for the infrasound of an evaporating cloud of water droplets in a dry environment. Different results are obtained for different constraints on the evolution of the droplet size distribution, which are commonly imposed by c-RAMS and other cloud models. The analytical results provide benchmarks for evaluating the ability of c-RAMS to simulate infrasound consistent with its theoretical foundation.

a. The pressure equation

Consider an atmosphere at rest that contains a spherical cloud of water droplets. The cloud is characterized by the liquid mixing ratio
i1558-8432-49-4-664-e1
in which x is the position vector relative to the center of the cloud, t is time, ϵ is the cloud radius, and Θ is the Heaviside step function. The value of Θ is zero or unity if its argument is negative or positive, respectively.

If the air is subsaturated, the water droplets evaporate. Evaporation cools the air and increases the vapor mixing ratio rυ(x, t) within the cloud. The combination of local cooling and gaseous mass production generates an outward propagating acoustic pulse. The following derives a formal expression for the pressure perturbation p′(x, t) associated with the pulse. The derivation is based on the thermo-mechanical core of c-RAMS. For simplicity, the ambient pressure p0, mass density ρ0, absolute temperature T0, and sound speed c0 ≡ (cpp0/cυρ0)1/2 are treated as constants. The symbols cp, cυ, and R represent the specific heat at constant pressure, the specific heat at constant volume, and the gas constant of dry air. In general, a prime or zero subscript denotes a perturbation or basic-state variable, respectively.

We assume that all perturbations are sufficiently weak to justify linearizing the compressible gas dynamics. The linearized mass continuity equation, including a source term due to evaporation, is given by
i1558-8432-49-4-664-e2
in which ρ is the mass density of the gaseous component of moist air and u is the velocity field. The linearized momentum equation is
i1558-8432-49-4-664-e3
assuming that gravitational effects and viscosity are unimportant over the time scale of interest.
The definition of virtual potential temperature,
i1558-8432-49-4-664-e4
here serves as the equation of state. Linearizing Eq. (4) yields
i1558-8432-49-4-664-e5
Neglecting diffusion by subgrid eddies and sedimentation, the heat equation in RAMS is approximated by the material conservation of ice–liquid potential temperature, defined by (Tripoli and Cotton 1981)
i1558-8432-49-4-664-e6
Here, Llv is the latent heat of vaporization (per unit mass), Liv is the latent heat of sublimation, ri is the mixing ratio of ice, and T is the absolute temperature of the air. Taking into consideration that T > 253 K, ri = 0, rυ ≪ 1, Llvrl/cpT ≪ 1, and rl = −rυ, the linearized heat equation reduces to
i1558-8432-49-4-664-e7
Taking the time derivative of Eq. (2), and eliminating u, ρ′, and θυ with Eqs. (3), (5), and (7) yields
i1558-8432-49-4-664-e8
Note that the factor in parentheses is typically negative and much greater than unity, meaning that evaporative cooling dominates the mass source of the acoustic emission. A formal solution to Eq. (8) in an infinite domain is given by
i1558-8432-49-4-664-e9
in which t − |xy|/c0 is the retarded time. The reader may consult appendix B for a generalization of Eq. (8) that accounts for ice.

b. Evaporation equations for two single-moment schemes

Neglecting the convective component of the material derivative, which is second order in the velocity and mixing ratio, the rate of change of water vapor is given by
i1558-8432-49-4-664-e10
in which r*c is the saturation mixing ratio at the surface of a cloud droplet. The bulk rate of evaporation is largely determined by the value of
i1558-8432-49-4-664-e11
in which ψ is the vapor diffusivity, N is the number concentration of cloud droplets, Dm is the mean droplet diameter, and S is the droplet shape parameter. Here, we have neglected a small correction to μ associated with the characteristic Reynolds number of the droplet motion. To close the evaporation equation [Eq. (10)], μ and r*c must be expressed as functions of rυ.
The first step of the μ closure is to consider the relationship between rl, N, and Dm. In RAMS, the probability distribution of the droplet diameter D is given by
i1558-8432-49-4-664-e12
in which ν is an adjustable parameter and Γ is the standard gamma function. The mass density of cloud droplets is given by , in which m is the mass of an individual droplet. By assumption, m = αDβ, in which α and β are empirical parameters. The mixing ratio of cloud water is thus
i1558-8432-49-4-664-e13
In a single-moment model, either N or Dm is specified as a fixed parameter of the hydrometeor distribution. An expression for the other parameter as a function of rυ in the evaporating cloud is obtained from Eq. (13) and the conservation of water mass (rl = −rυ). Substituting the result into Eq. (11) yields
i1558-8432-49-4-664-e14
in which rtrυ0 + rl0 and rυ0 is the initial vapor mixing ratio. Although the constant-N model (top) may seem more physical for the problem under consideration, the constant-Dm model (bottom) is often used for precipitating hydrometeors.
To determine r*c, we suppose that the net heat flux (conductive plus latent) at the surface of each cloud droplet is approximately zero. This leads to an equation of the form
i1558-8432-49-4-664-e15
in which Tc is the temperature at the surface of a droplet and κ is the thermal conductivity of air (e.g. Pruppacher and Klett 1997). In this section, we simplify Eq. (15) by assuming
i1558-8432-49-4-664-e16
which makes sense only for low-density clouds in dry environments. The condition in Eq. (16) permits setting rυ equal to zero on the left-hand side of Eq. (15). We further assume that the liquid water mass is sufficiently small for all other variables in the equation to stay nearly fixed over the course of evaporation. A standard analytical formula for Tc(r*c, p0) (e.g., Emanuel 1994; Walko et al. 1995) may then be used to convert Eq. (15) into a time-invariant equation for r*c.
Substituting Eq. (14) and a constant value of r*c into Eq. (10) yields an autonomous, first-order, ordinary differential equation (ODE) for rυ. Assuming again that the air remains sufficiently dry throughout the evaporation process [Eq. (16)], we replace r*crυ with r*c on the right-hand side. The solution of the ODE for the constant-N model is then
i1558-8432-49-4-664-e17
in which
i1558-8432-49-4-664-e18
Here (and below), we have assumed an initial vapor mixing ratio of rυ0 = 0, which corresponds to the numerical experiments of section 3. Note that evaporation ends (rυ = rl0) at t = γ−1. The solution of the ODE for the constant-Dm model is given by
i1558-8432-49-4-664-e19
in which
i1558-8432-49-4-664-e20
In this case, evaporation continues forever.

c. Solutions of the pressure equation

Given either of the above formulas for rυ, the integral equation for the pressure perturbation [Eq. (9)] is readily evaluated. Consider a spherical coordinate system whose origin is at the observation point, which is a distance x from the center of the cloud. Let ϖ, λ, and φ denote the radius, azimuth, and polar angle (measured from the axis connecting the origin and the center of the cloud). In this coordinate system, we may write ∂ttrυP(t)Θ(ϵy), in which y ≡ (x2 + ϖ2 − 2 cosφ)1/2. The specific form of P depends on the single-moment model that is used for the microphysics. Both the constant-N (top) and constant-Dm (bottom) forms are given below:
i1558-8432-49-4-664-e21
in which δ is the Dirac distribution. The top and bottom definitions of γ are given by Eqs. (18) and (20), respectively.
Continuing, let Bρ0(1.61 − Llv/cpT0)/4π. Then the integral equation [Eq. (9)] may be written
i1558-8432-49-4-664-e22
in which = t − ϖ/c0. Upon further reduction, we obtain
i1558-8432-49-4-664-e23
under the assumption that x > ϵ. What remains is to substitute (21) into (23) with t, and to perform the integration.
The solutions are conveniently expressed in terms of the following dimensionless variables:
i1558-8432-49-4-664-e24
For the case of evaporation with constant N we have
i1558-8432-49-4-664-e25
in which
i1558-8432-49-4-664-e26
and
i1558-8432-49-4-664-e27
The value of F1 for any τ is obtained by numerical quadrature. For the case of evaporation with constant Dm we have
i1558-8432-49-4-664-e28
in which
i1558-8432-49-4-664-e29
Section 3 discusses the basic properties of solutions (25) and (28), and the extent to which they agree with the infrasound that is simulated by c-RAMS.

3. C-RAMS simulations of infrasound generated by evaporation in a dry environment

The analytical pressure perturbations of section 2 are now compared with the infrasound simulated by c-RAMS when a cloud evaporates in a dry environment. We consider six numerical experiments with variable microphysics. The common experimental settings (of interest) are listed in Table 1.

A subset of the parameters in Table 1 pertains to the discretization. The model is configured to contain a fine inner grid (i) and a coarse outer grid (o). The inner horizontal grid spacing (dxi = dxo/3) is one-tenth of the cloud radius ϵ. The vertical grid spacing (dz = dxi) is uniform. Both the inner and outer grids have sufficient spatial resolutions to simulate 1-s (and longer) thermo-acoustic processes; the temporal resolutions (dti = dto/2 = 0.02 s) are also adequate. The horizontal widths of the inner and outer grids (Xi and Xo) are roughly 10 and 40 times ϵ, respectively. The lateral boundary conditions of the outer grid are set to allow free passage of acoustic waves (cf. Klemp and Wilhelmson 1978). The vertical boundaries at the ground and height Z are purely reflective; however, they are far enough removed from the cloud to prevent echoes from returning over the time period under consideration.

The basic state of the atmosphere is dry and isentropic. The vertical density and pressure gradients that would ordinarily appear, because of enforcement of hydrostatic balance, are eliminated by setting the gravitational acceleration g to a very small value (10−4 m s−2). At t = 0, a cloud is introduced by adding a uniform spherical distribution of rl [Eq. (1)] in the center of the domain, and adjusting the ice–liquid potential temperature [Eq. (6)] accordingly. In principle, either “cloud water” or “rain” (two distinct hydrometeor categories in RAMS) may be used to form the cloud mass. Here, we use a customized rain category, which has zero fall speed and an initial temperature of Tc = 281.2 K, which corresponds to a surface saturation mixing ratio of r*c = 6.8 × 10−3.

All six experiments use Marshall–Palmer distributions (ν = 1) for the droplet size. The distributions have three distinct initial conditions, summarized in rows 1–3 of Table 2. For each initial condition, the cloud is allowed to evaporate keeping either N or Dm fixed. Figure 1 shows the acoustic emission (diamonds) for each case at x = 424.3 m. The pressure perturbation is multiplied by to compensate for “geometric decay” with distance from the edge of the cloud.

The top row of Fig. 1 shows the variation of the acoustic emission with the number density N, when N is held fixed during evaporation. Decreasing N by two orders of magnitude severely damps the peak wave amplitude. Although reducing N has little influence on the width of the leading (negative) pulse, it dramatically widens the trailing (positive) wave. The bottom row shows the variation of the acoustic emission with the mean droplet diameter Dm, when Dm is held fixed during evaporation. Each plot is directly underneath the constant-N experiment with the same microscopic initial conditions (Table 2). Evidently, changing the single-moment scheme significantly modifies the shape of the trailing wave, but not the characteristic amplitude or time scale of the emission.

The basic properties of the acoustic emissions are readily explained. To begin with, the leading pulse is created by the initial shock of the evaporation rate over the entire cloud. Because of the finite propagation speed of acoustic signals, an observer senses this shock over a time period of 2ϵ/c0. The duration of the trailing wave is the evaporation time scale, which dilates with decreasing N or increasing Dm. Finally, since the amplitude of the acoustic source decreases with the evaporation rate, so must the emission attenuate with decreasing N or increasing Dm, for a given rl0 [see Eq. (14)].

The solid curves in each figure correspond to the integral on the right-hand side of Eq. (22), with P() given by the simulated time series of rυ. All solid curves very closely match the pressure waves that are generated by c-RAMS. In this sense, c-RAMS correctly simulates the infrasound of evaporation.2

The dotted curves in each figure correspond to the analytical approximations of section 2 [Eqs. (25) and (28)]. When Dm is held constant during evaporation, the approximations are excellent. When N is held constant during evaporation, a quantitative discrepancy is noticeable. The error appears primarily because our calculated value of Tc, based on Eq. (15) with rυ = 0, is roughly 1 K less than the time-averaged hydrometeor temperature (Tc) in c-RAMS. The associated 5% discrepancy in TTc is plausible, since the vapor mixing ratio grows with time, and the microphysics algorithm in c-RAMS does not strictly enforce Eq. (15) (Walko et al. 2000). The dash–dotted curves are the analytical approximations using Tc. For fixed N, this revision clearly improves the quantitative agreement between the analytical waves and the infrasound that is generated by c-RAMS. For fixed Dm, the revised curves are nearly indistinguishable from the data.

4. Infrasound generated by evaporation in a humid environment

This section examines the infrasound of an evaporating cloud in a humid environment. For analytical convenience, we restrict our analysis to the single-moment parameterization in which Dm is constant.

a. Basic theory

Factoring μ into two parts, we may rewrite the evaporation equation as follows:
i1558-8432-49-4-664-e30
in which
i1558-8432-49-4-664-e31
If the air is nearly saturated, the term r*crυ on the right-hand side of Eq. (30) can change over time by a large fraction of its initial value. We may express this term as a function of the variable rυ alone, by substituting the following approximations into Eq. (15):
i1558-8432-49-4-664-e32
in which r*0 and (∂r*/∂T)0 are the saturation mixing ratio and its temperature derivative at constant pressure, evaluated at T0 and p0. The new version of Eq. (15) is given by
i1558-8432-49-4-664-e33
in which
i1558-8432-49-4-664-e34
The reader may consult appendix C for further discussion of the key approximations used in deriving this result.
The solution of Eq. (30) is conveniently expressed in terms of the following two variables, related to the initial degree of subsaturation and the initial liquid mixing ratio:
i1558-8432-49-4-664-e35
Specifically, we have
i1558-8432-49-4-664-e36
in which
i1558-8432-49-4-664-e37
and
i1558-8432-49-4-664-e38
The value of γ can be positive or negative. When γ is positive (ζ0 < 1), the cloud fully evaporates over time. When γ is negative (ζ0 > 1), saturation occurs before the condensate is fully removed. In this case, the final value of the liquid mixing ratio is rlf = rl0ξ0/(1 + b). Note that ζ0 is the ratio of the initial liquid water mass (rl0) to the maximum that can evaporate (rl0rlf).
As before, we may use Eq. (23) to calculate the pressure perturbation outside of the cloud. Solving the integral yields
i1558-8432-49-4-664-e39
in which
i1558-8432-49-4-664-e40
The dimensionless variables , τ, γ̃, and were defined earlier in Eq. (24).
The wave form is similar to the acoustic pulse generated by evaporation in a dry environment. A negative minimum occurs at about τ = 0, whereas a positive maximum occurs at τ = 1. If γ̃ ≪ 1, one may derive the following expressions for the minimum (−) and maximum (+) values of the pressure perturbation:
i1558-8432-49-4-664-e41
Here, we have introduced the (approximate) initial cloud mass,
i1558-8432-49-4-664-e42
and the initial decay rate of the cloud mass,
i1558-8432-49-4-664-e43
Note that p+ depends on the cloud size (ϵ) only through Mc0, and is proportional to the square of Ê0. Because a cloud-size shock creates the leading wave, p varies with ϵ through both Mc0 and c0/ϵ, which is the regularized shock frequency. In contrast to p+, the variation of p with Ê0 is linear.
The decay rate (characteristic frequency) of the trailing wave corresponds to the “proper” evaporation rate, defined by
i1558-8432-49-4-664-e44
in which rl is the time-asymptotic liquid mixing ratio. If the air saturates without consuming the entire cloud, then rl is nonzero and E0 > Ê0. Figure 2 illustrates how E0 varies with relative humidity, liquid water mass, droplet size, and air temperature.3 In general, increasing the relatively humidity (rυ0/r*0) reduces the evaporation rate until ζ0 exceeds unity, after which E0 remains constant. For fixed relative humidity, the evaporation rate decays with the liquid water mass until ζ0 drops below unity. The evaporation rate also decays with increasing droplet size, as Dm−2 if β = 3. As a final remark on the matter, lowering the air temperature (with an adiabatic pressure drop) tends to decelerate evaporation for a given relative humidity, droplet size, and liquid water mass.

Figure 3 illustrates how the peak amplitude (p+) of the trailing wave of the infrasonic emission varies with relative humidity, liquid water mass, droplet size, and air temperature. In general, the numerical values correspond to pressure perturbations at x = 10 km from the center of a cloud of radius ϵ = 100 m. Sensibly, increasing the relative humidity to 100% damps the wave amplitude to zero. Decreasing the liquid water mass also leads to a weaker emission. Increasing the droplet size decreases the trailing wave amplitude asymptotically as Dm−4, assuming that β = 3. Lowering the air temperature (with an adiabatic pressure drop) also attenuates the wave, for a given relative humidity, droplet size, and liquid water mass.

Figure 4 illustrates how the trailing wave amplitude varies relative to the leading wave amplitude as the ambient water vapor and droplet size increase. Evidently, the magnitude of p+/p tends toward a constant value as the relative humidity tends toward 100%. From Eq. (41), we obtain the following approximation for this value:
i1558-8432-49-4-664-e45
Equation (41) also predicts the readily seen decay of p+/p with increasing droplet size (as Dm−2 if β = 3).

The paradigm of infrasound generated by the evaporation of an isolated cloud, suddenly introduced into a subsaturated environment, was conceived for the purpose of testing c-RAMS. It does not adequately represent evaporation in a realistic storm, which may result from turbulent mixing or falling hydrometeors.4 Nevertheless, the results shown here provide some basis for estimating the influence of evaporation (or condensation) on the infrasound in a storm simulation. We are primarily interested in the 0.1–10-Hz component of the infrasound, where severe storm signals are relatively strong and may contain detectable signatures of tornadoes (Bedard 2005). At 10 km from the source, the observed signals have wave amplitudes of the order 0.1 Pa (Bedard 2005; Schecter et al. 2008). The plots shown here loosely suggest that evaporation over 100-m-scale regions of high relative humidity (>90%) may noticeably affect these signals only if the cloud water has a mean droplet size of about 10 μm or less, or a liquid mixing ratio greater than 0.1 g kg−1. Otherwise, phase transitions are too slow and the acoustic emissions are too weak.

b. Comparison with c-RAMS

Figure 5 compares the analytical pressure perturbations of section 4a with the infrasound simulated by c-RAMS due to evaporation under humid conditions. For proper comparison, all three numerical experiments use the single-moment parameterization in which Dm is held fixed. Table 1 lists the common parameters of all experiments (note that Table 1 neglects very small corrections to the ambient mass density and sound speed, associated with finite relative humidity), whereas Table 2 (rows 4–6) lists the distinguishing parameters. The left and middle plots correspond to simulations with the same relative humidity and liquid mixing ratio, but different values of Dm. The middle and right plots correspond to simulations with the same Dm, but different relative humidities and liquid mixing ratios.

Regardless of the specific conditions, the infrasonic emissions simulated by c-RAMS (solid curves) compare favorably to the analytical predictions (dotted curves) that are given by Eq. (39). We conclude that c-RAMS properly simulates the infrasound of parameterized evaporation in a humid environment.

5. Conclusions

In this paper, we derived analytical formulas for the infrasound generated by evaporating clouds in dry and humid environments. The derivations were based on the dynamical core of c-RAMS and standard single-moment microphysics parameterizations. The theoretical development elucidated the fundamental mechanism by which (liquid–vapor) phase transitions produce infrasound in the model. Furthermore, it explained the potential variation of 0.1–10-Hz emissions with the selected microphysics parameterization.

More important is that the analytical solutions gave us a benchmark for evaluating the ability of c-RAMS to simulate infrasound consistent with theory. Successful comparisons between the analytical results and computational output verified the adequacy of the numerics, under both dry and humid conditions. A more general theoretical benchmark covering the infrasound of ice and mixed-phase hydrometeors was left for another study. Nevertheless, appendix B shows that c-RAMS correctly generates the infrasound of an equilibrating cloud of mixed-phase hydrometeors, under the assumption that the microphysics algorithm properly models the ice–liquid–vapor phase conversions.

In summary, we have demonstrated that c-RAMS is built upon a solid foundation for simulating the infrasound associated with phase transitions of moisture. Fine details of the acoustic power spectrum (between 0.1 and 10 Hz) may vary with the microphysics parameterization, because phase transition rates are influenced by the imposed constraints on hydrometeor distributions. In our view, the subtle variation of spectral details should not discourage researchers from using c-RAMS (or any other cloud model) to investigate the conditions under which tornado infrasound may dominate the infrasound of generic moist processes within a storm. However, microphysics sensitivity tests may be necessary to establish definitive conclusions.

Acknowledgments

The authors thank Stephen Saleeby for a helpful discussion regarding the microphysics parameterizations in RAMS. The authors also thank three anonymous reviewers for their constructive comments on the original manuscript. This work was supported by NSF Grant ATM-0832320.

REFERENCES

  • Akhalkatsi, M., and G. Gogoberidze, 2009: Infrasound generation by tornadic supercell storms. Quart. J. Roy. Meteor. Soc., 135 , 935940.

    • Search Google Scholar
    • Export Citation
  • Bedard Jr., A. J., 2005: Low-frequency atmospheric acoustic energy associated with vortices produced by thunderstorms. Mon. Wea. Rev., 133 , 241263.

    • Search Google Scholar
    • Export Citation
  • Bedard Jr., A. J., B. W. Bartram, A. N. Keane, D. C. Welsh, and R. T. Nishiyama, 2004: The infrasound network (ISNet): Background, design details, and display capability as an 88D adjunct tornado detection tool. Preprints, 22nd Conf. on Severe Local Storms, Hyannis, MA, Amer. Meteor. Soc., 1.1. [Available online at http://ams.confex.com/ams/pdfpapers/816561.pdf].

    • Search Google Scholar
    • Export Citation
  • Cotton, W. R., and Coauthors, 2003: RAMS 2001: Current status and future directions. Meteor. Atmos. Phys., 82 , 529.

  • Emanuel, K. A., 1994: Atmospheric Convection. Oxford University Press, 580 pp.

  • Georges, T. M., 1976: Infrasound from convective storms. Part II: A critique of source candidates. NOAA Tech. Rep. ERL 380-WPL 49, 59 pp. [Available from the National Technical Information Service, 5285 Port Royal Rd., Springfield, VA 22161].

    • Search Google Scholar
    • Export Citation
  • Klemp, J. B., and R. B. Wilhelmson, 1978: The simulation of three-dimensional convective storm dynamics. J. Atmos. Sci., 35 , 10701093.

    • Search Google Scholar
    • Export Citation
  • Meyers, M. P., R. L. Walko, J. Y. Harrington, and W. R. Cotton, 1997: New RAMS cloud microphysics parameterization. Part II: The two-moment scheme. Atmos. Res., 45 , 339.

    • Search Google Scholar
    • Export Citation
  • Nicholls, M. E., and R. A. Pielke Sr., 1994a: Thermal compression waves I: Total energy transfer. Quart. J. Roy. Meteor. Soc., 120 , 305332.

    • Search Google Scholar
    • Export Citation
  • Nicholls, M. E., and R. A. Pielke Sr., 1994b: Thermal compression waves II: Mass adjustment and vertical transfer of total energy. Quart. J. Roy. Meteor. Soc., 120 , 333359.

    • Search Google Scholar
    • Export Citation
  • Nicholls, M. E., and R. A. Pielke Sr., 2000: Thermally induced compression waves and gravity waves generated by convective storms. J. Atmos. Sci., 57 , 32513271.

    • Search Google Scholar
    • Export Citation
  • Passner, J. E., and J. M. Noble, 2006: Acoustic energy measured in severe storms during a field study in June 2003. Army Research Laboratory Tech. Rep. ARL-TR-3749, 42 pp.

    • Search Google Scholar
    • Export Citation
  • Pruppacher, H. R., and J. D. Klett, 1997: Microphysics of Clouds and Precipitation. Kluwer Academic, 954 pp.

  • Saleeby, S. M., and W. R. Cotton, 2004: A large-droplet mode and prognostic number concentration of cloud droplets in the Colorado State University Regional Atmospheric Modeling System (RAMS). Part I: Module descriptions and supercell test simulations. J. Appl. Meteor., 43 , 182195.

    • Search Google Scholar
    • Export Citation
  • Schecter, D. A., M. E. Nicholls, J. Persing, A. J. Bedard Jr., and R. A. Pielke Sr., 2008: Infrasound emitted by tornado-like vortices: Basic theory and a numerical comparison to the acoustic radiation of a single-cell thunderstorm. J. Atmos. Sci., 65 , 685713.

    • Search Google Scholar
    • Export Citation
  • Szoke, E. J., A. J. Bedard Jr., E. Thaler, and R. Glancy, 2004: A comparison of ISNet data with radar data for tornadic and potentially tornadic storms in northeast Colorado. Preprints, 22nd Conf. on Severe Local Storms, Hyannis, MA, Amer. Meteor. Soc., 1.2. [Available online at http://ams.confex.com/ams/pdfpapers/81466.pdf].

    • Search Google Scholar
    • Export Citation
  • Tripoli, G. J., and W. R. Cotton, 1981: The use of ice-liquid potential temperature as a thermodynamic variable in deep atmospheric models. Mon. Wea. Rev., 109 , 10941102.

    • Search Google Scholar
    • Export Citation
  • Walko, R. L., W. R. Cotton, M. P. Meyers, and J. Y. Harrington, 1995: New RAMS cloud microphysics parameterization. Part 1: The single-moment scheme. Atmos. Res., 38 , 2962.

    • Search Google Scholar
    • Export Citation
  • Walko, R. L., W. R. Cotton, G. Feingold, and B. Stevens, 2000: Efficient computation of vapor and heat diffusion between hydrometeors in a numerical model. Atmos. Res., 53 , 171183.

    • Search Google Scholar
    • Export Citation

APPENDIX A

Customization of RAMS

In RAMS, the mass continuity equation is replaced with a prognostic equation for the Exner function, defined by Π ≡ cp(p/p0)R/cp. The complete form of this equation (including a term connected to mass exchange between air and hydrometeors) is given by
i1558-8432-49-4-664-ea1
in which Π′ is the perturbation Exner function and c2RΠθυ/cυ. All terms in Eq. (A1) are kept in the customized version of RAMS that is used here to study atmospheric acoustics (cf. Nicholls and Pielke 2000). The standard version of RAMS neglects the entire right-hand side.

Note that while perhaps unnecessary, the sedimentation routine was commented out of the code for this study. Furthermore, the macroscopic diffusion coefficients (connected to subgrid eddies) were given negligible values.

APPENDIX B

Linearized Pressure Equation with Ice

The linearized heat equation [Eq. (7)] is readily generalized to include ice and mixed-phase hydrometeors as follows:
i1558-8432-49-4-664-eb1
Neglecting diffusion by subgrid eddies and sedimentation, conservation of water mass implies that
i1558-8432-49-4-664-eb2
Taking the time derivative of Eq. (2), and eliminating u, ρ′, θυ, and rl with Eqs. (3), (5), (B1), and (B2) yields
i1558-8432-49-4-664-eb3
Given the time series of rυ and ri, the solution of (B3) outside of a uniform cloud reduces to an integral similar to the right-hand side of Eq. (23).

Figure B1 verifies that the integral for p′ agrees with the infrasound generated by c-RAMS, when a uniform sphere of hail (ϵ = 200 m) is suddenly introduced into a warm (T0 = 300 K) and saturated (rυ0 = r*0) environment. The experiment employs a single-moment parameterization in which the mean diameter (Dm) of the hail distribution is held constant at 1 cm, and ν = 2. The initial value of the hail mixing ratio is 1.14 g kg−1, and the initial liquid mass fraction is zero. Over time, the ice melts under the actions of heat and vapor diffusion, as explained in standard textbooks (e.g., Pruppacher and Klett 1997). Because the mean surface-to-volume ratio of hail is relatively small, phase transitions occur slowly. Consequently, the only significant infrasound is produced by the initial shock. The dotted curves show the individual contributions to the “shock wave” computed from the output vapor and ice (top and bottom) terms on the right-hand side of (B3). The solid curve is the sum of both contributions, which matches the pressure perturbation generated by c-RAMS (diamonds).

Note that the time step (dti) for the hail simulation is 2 ms, and the standard value of Liv = 2.834 × 106 J kg−1 is used to calculate the theoretical infrasound.

APPENDIX C

Approximations in the Evaporation Equation under Humid Conditions

Equation (32) provides an approximation for the air temperature T that is derived from the following reformulated heat equation [Eq. (7)]:
i1558-8432-49-4-664-ec1
For a humid environment, we may assume that the evaporation rate E is sufficiently small for the acoustic wavelength c0/E to greatly exceed the length scale l of the source region. Under this condition, Eq. (8) suggests that p′ ∼ l2E2ρ0Llvrυ/cpT0. Accordingly, the ratio of the p′ term over the rυ term in Eq. (C1) is of the order (l2E2/c02)(R/cυ) ≪ 1. For this reason, Eq. (32) neglects the partial variation of T ′ with p′.
Equation (32) also provides an approximation for the hydrometeor temperature Tc. This formula assumes that dTcTcT0 is very small. From the relations
i1558-8432-49-4-664-ec2
we obtain
i1558-8432-49-4-664-ec3
in which Rυ (R) is the gas constant of water vapor (dry air) and e*(T) ≪ p denotes the saturation vapor pressure (e.g., Emanuel 1994). Using Eq. (33) to evaluate dr*cr*cr*0, and using our previous estimate of dpp′, one may show that neglecting the term proportional to dp in Eq. (C3) is typically consistent.

Fig. 1.
Fig. 1.

The infrasound of an evaporating cloud in a dry environment, with six different microphysics parameterizations. (top) Experiments with fixed N (as indicated) and time-dependent Dm. (bottom) Experiments with fixed Dm (as indicated) and time-dependent N. Each column has the same initial values for N and Dm. Diamonds correspond to the infrasound simulated by c-RAMS (only a fraction of the data is shown for clarity). The solid curve corresponds to the infrasound that is theoretically generated by the output time series of rυ. The dotted and dash–dotted curves correspond to two variants of an analytical model that predicts both rυ and p′ (see text). The pressure perturbations are multiplied by x/ϵ. Note that one unit of τ ≡ (c0tx)/ϵ corresponds to a 0.58-s change of t.

Citation: Journal of Applied Meteorology and Climatology 49, 4; 10.1175/2009JAMC2226.1

Fig. 2.
Fig. 2.

(top) The characteristic evaporation rate (E0) vs relative humidity (RH), for several values of the liquid mixing ratio rl0. All curves are calculated with Dm = 10 μm, T0 = 300 K, and p0 = 105 Pa. (bottom) Evaporation rate E0 vs RH for several values of the mean droplet diameter Dm. All curves are calculated with rl0 = 4.6 × 10−4, T0 = 300 K, and p0 = 105 Pa, except for the lower solid (5 μm) curve, in which T0 = 275 K and p0 = 73 757 Pa.

Citation: Journal of Applied Meteorology and Climatology 49, 4; 10.1175/2009JAMC2226.1

Fig. 3.
Fig. 3.

(top) The pressure peak of the trailing wave at x = 10 km vs RH, for several values of the liquid mixing ratio rl0. All curves are calculated with a cloud radius of ϵ = 100 m, except for the lower dashed (rl0 = 0.01r*0) curve, for which ϵ = 10 m. Furthermore, all curves are calculated with Dm = 5 μm, T0 = 300 K, and p0 = 105 Pa. (bottom) The pressure peak of the trailing wave at x = 10 km vs RH for several values of the mean droplet diameter Dm. All curves are calculated with ϵ = 100 m, rl0 = 2.28 × 10−3, T0 = 300 K, and p0 = 105 Pa, except for the lower dashed (10 μm) curve, for which T0 = 275 K and p0 = 73 757 Pa.

Citation: Journal of Applied Meteorology and Climatology 49, 4; 10.1175/2009JAMC2226.1

Fig. 4.
Fig. 4.

Variation of the wave asymmetry (|p+/p|) with RH and mean droplet size. The air temperature and the liquid mixing ratio (relative to r*0) are specified on the plot. The cloud radius ϵ is 100 m. The diamonds on the right axis mark the approximate saturation limits for each curve, given by Eq. (45).

Citation: Journal of Applied Meteorology and Climatology 49, 4; 10.1175/2009JAMC2226.1

Fig. 5.
Fig. 5.

Infrasound generated by evaporating clouds under humid conditions. The solid and dotted curves correspond to c-RAMS and theoretical predictions, respectively. The distinguishing characteristics of each experiment (and the common temperature) are printed in the lower-right corner of each plot. The pressure perturbations are multiplied by x/ϵ.

Citation: Journal of Applied Meteorology and Climatology 49, 4; 10.1175/2009JAMC2226.1

i1558-8432-49-4-664-fb01

Fig. B1. Infrasound generated by suddenly introducing hail into warm, saturated air. The dotted curves are the theoretical waves produced by the vapor and ice source terms on the right-hand side of Eq. (B3). The solid curve is the combined wave, and the diamonds represent the infrasound simulated by c-RAMS. As usual, the pressure perturbation is multiplied by x/ϵ.

Citation: Journal of Applied Meteorology and Climatology 49, 4; 10.1175/2009JAMC2226.1

Table 1.

Common experimental parameters, defined in the text.

Table 1.
Table 2.

Distinct initial conditions for experiments in dry (rows 1–3) and humid (rows 4–6) environments. All parameters are defined in sections 2 and 4.

Table 2.

1

The term “infrasound” refers to sound waves at frequencies less than the lower limit of unimpaired human hearing, which is roughly 20 Hz.

2

A similar test has verified the simulated infrasound of evaporation governed by the two-moment microphysics option of c-RAMS (Meyers et al. 1997). In this case, the trailing wave exhibits gradual decay at late times, qualitatively similar to the infrasound generated with fixed Dm.

3

Table 1 gives the unspecified parameters that are required to evaluate most curves in Figs. 2 –4, for which T0 = 300 K. For the exceptional cases where T0 = 275 K, the pressure is reduced (dry adiabatically) to p0 = 105 (275/300)cp/R = 73 757 Pa; furthermore, the values of c0, ρ0, ψ, κ, r*0, and (∂r*/∂T)0 are adjusted according to standard formulas (e.g., Pruppacher and Klett 1997; Emanuel 1994).

4

In a practical storm simulation, the subgrid turbulence parameterization may have a switch that turns on when the local Richardson number is sufficiently small. Once activated, the model could rapidly mix finescale inhomogeneities of the entropy and moisture variables. Conceivably such mixing could (artificially) trigger the type of sudden phase transition considered here.

Save
  • Akhalkatsi, M., and G. Gogoberidze, 2009: Infrasound generation by tornadic supercell storms. Quart. J. Roy. Meteor. Soc., 135 , 935940.

    • Search Google Scholar
    • Export Citation
  • Bedard Jr., A. J., 2005: Low-frequency atmospheric acoustic energy associated with vortices produced by thunderstorms. Mon. Wea. Rev., 133 , 241263.

    • Search Google Scholar
    • Export Citation
  • Bedard Jr., A. J., B. W. Bartram, A. N. Keane, D. C. Welsh, and R. T. Nishiyama, 2004: The infrasound network (ISNet): Background, design details, and display capability as an 88D adjunct tornado detection tool. Preprints, 22nd Conf. on Severe Local Storms, Hyannis, MA, Amer. Meteor. Soc., 1.1. [Available online at http://ams.confex.com/ams/pdfpapers/816561.pdf].

    • Search Google Scholar
    • Export Citation
  • Cotton, W. R., and Coauthors, 2003: RAMS 2001: Current status and future directions. Meteor. Atmos. Phys., 82 , 529.

  • Emanuel, K. A., 1994: Atmospheric Convection. Oxford University Press, 580 pp.

  • Georges, T. M., 1976: Infrasound from convective storms. Part II: A critique of source candidates. NOAA Tech. Rep. ERL 380-WPL 49, 59 pp. [Available from the National Technical Information Service, 5285 Port Royal Rd., Springfield, VA 22161].

    • Search Google Scholar
    • Export Citation
  • Klemp, J. B., and R. B. Wilhelmson, 1978: The simulation of three-dimensional convective storm dynamics. J. Atmos. Sci., 35 , 10701093.

    • Search Google Scholar
    • Export Citation
  • Meyers, M. P., R. L. Walko, J. Y. Harrington, and W. R. Cotton, 1997: New RAMS cloud microphysics parameterization. Part II: The two-moment scheme. Atmos. Res., 45 , 339.

    • Search Google Scholar
    • Export Citation
  • Nicholls, M. E., and R. A. Pielke Sr., 1994a: Thermal compression waves I: Total energy transfer. Quart. J. Roy. Meteor. Soc., 120 , 305332.

    • Search Google Scholar
    • Export Citation
  • Nicholls, M. E., and R. A. Pielke Sr., 1994b: Thermal compression waves II: Mass adjustment and vertical transfer of total energy. Quart. J. Roy. Meteor. Soc., 120 , 333359.

    • Search Google Scholar
    • Export Citation
  • Nicholls, M. E., and R. A. Pielke Sr., 2000: Thermally induced compression waves and gravity waves generated by convective storms. J. Atmos. Sci., 57 , 32513271.

    • Search Google Scholar
    • Export Citation
  • Passner, J. E., and J. M. Noble, 2006: Acoustic energy measured in severe storms during a field study in June 2003. Army Research Laboratory Tech. Rep. ARL-TR-3749, 42 pp.

    • Search Google Scholar
    • Export Citation
  • Pruppacher, H. R., and J. D. Klett, 1997: Microphysics of Clouds and Precipitation. Kluwer Academic, 954 pp.

  • Saleeby, S. M., and W. R. Cotton, 2004: A large-droplet mode and prognostic number concentration of cloud droplets in the Colorado State University Regional Atmospheric Modeling System (RAMS). Part I: Module descriptions and supercell test simulations. J. Appl. Meteor., 43 , 182195.

    • Search Google Scholar
    • Export Citation
  • Schecter, D. A., M. E. Nicholls, J. Persing, A. J. Bedard Jr., and R. A. Pielke Sr., 2008: Infrasound emitted by tornado-like vortices: Basic theory and a numerical comparison to the acoustic radiation of a single-cell thunderstorm. J. Atmos. Sci., 65 , 685713.

    • Search Google Scholar
    • Export Citation
  • Szoke, E. J., A. J. Bedard Jr., E. Thaler, and R. Glancy, 2004: A comparison of ISNet data with radar data for tornadic and potentially tornadic storms in northeast Colorado. Preprints, 22nd Conf. on Severe Local Storms, Hyannis, MA, Amer. Meteor. Soc., 1.2. [Available online at http://ams.confex.com/ams/pdfpapers/81466.pdf].

    • Search Google Scholar
    • Export Citation
  • Tripoli, G. J., and W. R. Cotton, 1981: The use of ice-liquid potential temperature as a thermodynamic variable in deep atmospheric models. Mon. Wea. Rev., 109 , 10941102.

    • Search Google Scholar
    • Export Citation
  • Walko, R. L., W. R. Cotton, M. P. Meyers, and J. Y. Harrington, 1995: New RAMS cloud microphysics parameterization. Part 1: The single-moment scheme. Atmos. Res., 38 , 2962.

    • Search Google Scholar
    • Export Citation
  • Walko, R. L., W. R. Cotton, G. Feingold, and B. Stevens, 2000: Efficient computation of vapor and heat diffusion between hydrometeors in a numerical model. Atmos. Res., 53 , 171183.

    • Search Google Scholar
    • Export Citation
  • Fig. 1.

    The infrasound of an evaporating cloud in a dry environment, with six different microphysics parameterizations. (top) Experiments with fixed N (as indicated) and time-dependent Dm. (bottom) Experiments with fixed Dm (as indicated) and time-dependent N. Each column has the same initial values for N and Dm. Diamonds correspond to the infrasound simulated by c-RAMS (only a fraction of the data is shown for clarity). The solid curve corresponds to the infrasound that is theoretically generated by the output time series of rυ. The dotted and dash–dotted curves correspond to two variants of an analytical model that predicts both rυ and p′ (see text). The pressure perturbations are multiplied by x/ϵ. Note that one unit of τ ≡ (c0tx)/ϵ corresponds to a 0.58-s change of t.

  • Fig. 2.

    (top) The characteristic evaporation rate (E0) vs relative humidity (RH), for several values of the liquid mixing ratio rl0. All curves are calculated with Dm = 10 μm, T0 = 300 K, and p0 = 105 Pa. (bottom) Evaporation rate E0 vs RH for several values of the mean droplet diameter Dm. All curves are calculated with rl0 = 4.6 × 10−4, T0 = 300 K, and p0 = 105 Pa, except for the lower solid (5 μm) curve, in which T0 = 275 K and p0 = 73 757 Pa.

  • Fig. 3.

    (top) The pressure peak of the trailing wave at x = 10 km vs RH, for several values of the liquid mixing ratio rl0. All curves are calculated with a cloud radius of ϵ = 100 m, except for the lower dashed (rl0 = 0.01r*0) curve, for which ϵ = 10 m. Furthermore, all curves are calculated with Dm = 5 μm, T0 = 300 K, and p0 = 105 Pa. (bottom) The pressure peak of the trailing wave at x = 10 km vs RH for several values of the mean droplet diameter Dm. All curves are calculated with ϵ = 100 m, rl0 = 2.28 × 10−3, T0 = 300 K, and p0 = 105 Pa, except for the lower dashed (10 μm) curve, for which T0 = 275 K and p0 = 73 757 Pa.

  • Fig. 4.

    Variation of the wave asymmetry (|p+/p|) with RH and mean droplet size. The air temperature and the liquid mixing ratio (relative to r*0) are specified on the plot. The cloud radius ϵ is 100 m. The diamonds on the right axis mark the approximate saturation limits for each curve, given by Eq. (45).

  • Fig. 5.

    Infrasound generated by evaporating clouds under humid conditions. The solid and dotted curves correspond to c-RAMS and theoretical predictions, respectively. The distinguishing characteristics of each experiment (and the common temperature) are printed in the lower-right corner of each plot. The pressure perturbations are multiplied by x/ϵ.

  • Fig. B1. Infrasound generated by suddenly introducing hail into warm, saturated air. The dotted curves are the theoretical waves produced by the vapor and ice source terms on the right-hand side of Eq. (B3). The solid curve is the combined wave, and the diamonds represent the infrasound simulated by c-RAMS. As usual, the pressure perturbation is multiplied by x/ϵ.

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 275 60 5
PDF Downloads 118 38 3