a method for adaptive habit prediction in bulk microphysical...

16
A Method for Adaptive Habit Prediction in Bulk Microphysical Models. Part I: Theoretical Development JERRY Y. HARRINGTON AND KARA SULIA Department of Meteorology, The Pennsylvania State University, University Park, Pennsylvania HUGH MORRISON National Center for Atmospheric Research, Boulder, Colorado (Manuscript received 8 February 2012, in final form 28 August 2012) ABSTRACT Bulk microphysical schemes use the capacitance model for ice vapor growth in combination with mass–size relationships to determine the evolution of ice water content (IWC) and ice particle maximum dimension in time. These approaches are limited since a single axis length is used, the aspect ratio is usually held constant and mass–size relations have many available coefficients for similar ice types. Fixing the crystal aspect ratio severs the nonlinear link between aspect ratio changes and increased growth rates that occur during crystal growth. A method is presented here for predicting two crystal axes and the crystal aspect ratio in bulk models. Evolution of the ice mass mixing ratio is tied to the evolution of two axis length mixing ratios through the use of a historical axis ratio parameter containing memory of crystal shape. This parameter links the distributions of the two axes, allowing characterization of particle lengths using a single distribution. The method uses four prognostic variables: the mass and number mixing ratios, and two axis length mixing ratios. Development of the method is presented, with testing described in Part II. 1. Introduction The growth of ice crystals in nature is a complex phenomenon: atmospheric ice attains a variety of shapes that vary among relatively simple hexagonal plates and columns, dendritic structures, rosettes, and often irreg- ular crystals that may be polycrystalline in form. Though crystal shapes vary, they can be characterized by two axis lengths, a and c, and their primary shapes by the aspect ratio (f 5 c/a). For hexagonal prisms, the a di- mension is half the distance across the basal (hexagonal) face, whereas the c dimension is half the length of the prism (rectangular) face. Less ideal crystals are characterized in terms of their maximum and minimum dimensions: for flat, platelike particles a is half the maximum dimension (referred to here as the maximum semidimension) and c is half the minimum dimension, with the opposite being the case for columnlike particles. Traditionally, the primary habit is characterized by as- pect ratio variation with temperature, the classical pic- ture of which is based on laboratory and in situ data and is arguably most valid for T .2228C (e.g., Pruppacher and Klett 1997; Fukuta and Takahashi 1999; Bailey and Hallett 2009). These data suggest platelike (f , 1) crystals for temperatures in the ranges of 218 to 238C and 2108 to 2228C, whereas columnlike crystals (f . 1) occur in the ranges of 238 to 298C and lower than 2228C. More recent data suggest important modifications to the habit diagram at temperatures below about 2228C (e.g., Bailey and Hallett 2009), specifically that plates tend to occur more often and that columns are rare. Some mea- surements support this (e.g., Bailey and Hallett 2002, 2004); however, these data are at odds with other lab- oratory studies of ice crystal formation and growth (e.g., Libbrecht 2003). The secondary habit of ice crystals depends on the degree of supersaturation, and this clas- sification is most valid for T .2228C. Laboratory mea- surements generally show that lower saturation states produce compact hexagonal crystals depending in part on nucleation. As the saturation increases toward liq- uid, crystals tend to hollow: columns become needlelike Corresponding author address: Jerry Y. Harrington, 503 Walker Building, Department of Meteorology, The Pennsylvania State University, University Park, PA 16802. E-mail: [email protected] VOLUME 70 JOURNAL OF THE ATMOSPHERIC SCIENCES FEBRUARY 2013 DOI: 10.1175/JAS-D-12-040.1 Ó 2013 American Meteorological Society 349

Upload: others

Post on 14-Feb-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • A Method for Adaptive Habit Prediction in Bulk Microphysical Models.Part I: Theoretical Development

    JERRY Y. HARRINGTON AND KARA SULIA

    Department of Meteorology, The Pennsylvania State University, University Park, Pennsylvania

    HUGH MORRISON

    National Center for Atmospheric Research, Boulder, Colorado

    (Manuscript received 8 February 2012, in final form 28 August 2012)

    ABSTRACT

    Bulk microphysical schemes use the capacitancemodel for ice vapor growth in combination with mass–size

    relationships to determine the evolution of ice water content (IWC) and ice particle maximum dimension in

    time. These approaches are limited since a single axis length is used, the aspect ratio is usually held constant

    and mass–size relations have many available coefficients for similar ice types. Fixing the crystal aspect ratio

    severs the nonlinear link between aspect ratio changes and increased growth rates that occur during crystal

    growth. Amethod is presented here for predicting two crystal axes and the crystal aspect ratio in bulk models.

    Evolution of the ice mass mixing ratio is tied to the evolution of two axis length mixing ratios through the use

    of a historical axis ratio parameter containing memory of crystal shape. This parameter links the distributions

    of the two axes, allowing characterization of particle lengths using a single distribution. The method uses four

    prognostic variables: the mass and number mixing ratios, and two axis length mixing ratios. Development of

    the method is presented, with testing described in Part II.

    1. Introduction

    The growth of ice crystals in nature is a complex

    phenomenon: atmospheric ice attains a variety of shapes

    that vary among relatively simple hexagonal plates and

    columns, dendritic structures, rosettes, and often irreg-

    ular crystals that may be polycrystalline in form. Though

    crystal shapes vary, they can be characterized by two

    axis lengths, a and c, and their primary shapes by the

    aspect ratio (f 5 c/a). For hexagonal prisms, the a di-mension is half the distance across the basal (hexagonal)

    face, whereas the c dimension is half the length of

    the prism (rectangular) face. Less ideal crystals are

    characterized in terms of their maximum and minimum

    dimensions: for flat, platelike particles a is half the

    maximum dimension (referred to here as the maximum

    semidimension) and c is half the minimum dimension,

    with the opposite being the case for columnlike particles.

    Traditionally, the primary habit is characterized by as-

    pect ratio variation with temperature, the classical pic-

    ture of which is based on laboratory and in situ data and

    is arguably most valid for T . 2228C (e.g., Pruppacherand Klett 1997; Fukuta and Takahashi 1999; Bailey and

    Hallett 2009). These data suggest platelike (f , 1)crystals for temperatures in the ranges of 218 to 238Cand2108 to2228C, whereas columnlike crystals (f. 1)occur in the ranges of238 to298Cand lower than2228C.More recent data suggest important modifications to the

    habit diagram at temperatures below about 2228C (e.g.,Bailey and Hallett 2009), specifically that plates tend to

    occur more often and that columns are rare. Some mea-

    surements support this (e.g., Bailey and Hallett 2002,

    2004); however, these data are at odds with other lab-

    oratory studies of ice crystal formation and growth

    (e.g., Libbrecht 2003). The secondary habit of ice crystals

    depends on the degree of supersaturation, and this clas-

    sification is most valid for T . 2228C. Laboratory mea-surements generally show that lower saturation states

    produce compact hexagonal crystals depending in part

    on nucleation. As the saturation increases toward liq-

    uid, crystals tend to hollow: columns become needlelike

    Corresponding author address: Jerry Y. Harrington, 503 Walker

    Building, Department of Meteorology, The Pennsylvania State

    University, University Park, PA 16802.

    E-mail: [email protected]

    VOLUME 70 JOURNAL OF THE ATMOSPHER IC SC I ENCE S FEBRUARY 2013

    DOI: 10.1175/JAS-D-12-040.1

    � 2013 American Meteorological Society 349

  • near 268C, dendritic crystals appear around 2158C,with rosette and polycrystals forming at T , 2228C.It is presently impossible for cloud models to capture

    the variety of ice crystal masses, shapes, sizes, and dis-

    tributions that occur in real clouds. Nevertheless, all

    numerical models need some way to relate the mass of

    crystals to their shapes and sizes. Prior attempts have

    nearly universally involved using either equivalent vol-

    ume (or diameter) spheres (e.g., Lin et al. 1983; Reisner

    et al. 1998; Thompson et al. 2004) or mass–size re-

    lationships (e.g., Mitchell et al. 1990; Harrington et al.

    1995; Meyers et al. 1997; Woods et al. 2007; Thompson

    et al. 2008;Morrison andGrabowski 2008, 2010), though

    exceptions exist (e.g., Hashino and Tripoli 2007). These

    methods suffer from a number of deficiencies (Sulia and

    Harrington 2011, hereafter SH11). In comparison to

    wind tunnel data, the use of equivalent spheres un-

    derestimates ice growth at every temperature except

    around 2108 and 2228C where growth is nearly iso-metric (e.g., Fukuta and Takahashi 1999). Mass–size

    relations allow the prediction of only a single size, and

    the coefficients in the mass–size power laws are derived

    from in situ data and are therefore tied to the observed

    particle growth histories. These methods are not linked

    to the growth mechanisms that determine the evolution

    of the two primary crystal axes (a and c) and therefore

    cannot evolve crystals in a natural fashion.

    Bulk models are limited because they attempt to

    predict the evolution of only a few moments of the size

    spectrum, the shape of which is assumed a priori. Typi-

    cally, only the ice mass mixing ratio is predicted in single-

    moment schemes, whereas in two-moment schemes

    the number mixing ratio is predicted as well. Moreover,

    most current bulk microphysical methods artificially

    categorize particles based on predefined classes and

    then use transfer functions to move particles among

    classes as they evolve. For instance, Woods et al. (2007)

    predicts numerous ice classes as a way to deal with the

    problem of habit evolution, but such methods are com-

    putationally costly because of the increased number of

    prognostic variables. In the method of Harrington et al.

    (1995), transfer functions across a somewhat arbitrary

    size boundary are used to evolve smaller pristine ice

    class particles into a larger snow class. A more recent

    trend in microphysical modeling is to predict different

    ice particle properties for a given ice class. For instance,

    the method of Morrison and Grabowski (2008) predicts

    a rime mass fraction for ice crystals that avoids the tra-

    ditional approach of transferring ice to a separate graupel

    class, and Hashino and Tripoli (2007) use the Chen

    and Lamb (1994) approach for predicting habits in a

    hybrid-bin scheme. Such methods allow the use of fewer

    ice classes and provide for a more natural evolution of

    the ice particles based on their respective growth histo-

    ries. To evolve crystal habit, a way to relate the evo-

    lution of the a- and c-axis distributions is required, and

    this has been a significant hurdle for numerical models

    to surmount.

    The approach developed here predicts the evolution

    of ice particle habit by tracking the evolution of ice as-

    pect ratio following Chen and Lamb (1994, hereafter

    CL94) and SH11. Similarities exist in the approach to

    that presented in Hashino and Tripoli (2007), with the

    exception that the method presented here is derived

    for a bulk cloud model using four prognostic variables.

    The method provides a way to relate the a- and c-axis

    evolution through the use of laboratory-determined

    parameters, making the method amenable to improve-

    ments through new measurements. The approach elimi-

    nates thresholds between habits and follows the recent

    paradigm of predicting particle properties instead of

    multiple particle classes. The parameterization is de-

    veloped here with testing of the method described in

    Harrington et al. (2013, hereafter Part II).

    2. Single-particle evolution

    The Fickian-distribution method

    The bulk habit method is rooted in the work of CL94

    briefly described here, but a full discussion can be found

    in the original work (CL94), and an exposition of the

    physical implications in SH11. Fundamentally, the CL94

    method combines two predictive equations. The first is

    vapor diffusional mass growth that follows the classical

    capacitance model (Pruppacher and Klett 1997):

    dm

    dt5 4pC(c, a)Gi(T ,P)su,i 5 4pLfs(f)Gi(T,P)su,i

    5 4prfsr(f, r)Gi(T,P)su,i , (1)

    where C is the capacitance of the ice particle, Gi is

    a function that combines the effects of thermal and

    vapor diffusion processes [Lamb and Verlinde 2011,

    their Eq. (8.41), p. 343], P is the pressure, T is the

    temperature, and su,i is the ice supersaturation. Be-

    cause C depends on two axis lengths, it is difficult to

    integrate Eq. (1) in time. Most modeling applications

    use the second form to the right in which L is the

    maximum semidimension, and fs(f) 5 C/L is a shapefactor that depends only on the aspect ratio f. In this

    work, Eq. (1) is integrated over a time step using an

    equivalent volume sphere of radius r given by the

    rightmost equation and is used to avoid the numerical

    problem of evolving both a and c simultaneously.

    Evolving the mass in this fashion requires holding the

    350 JOURNAL OF THE ATMOSPHER IC SC IENCES VOLUME 70

  • shape factor, fsr 5C(c, a)/r, constant, which is accurateover short time steps (,10 s; see SH11).A major limitation of the traditional capacitance

    model, as SH11 and others (e.g., Nelson 1994, p. 85)

    point out, is that the theory produces a constant aspect

    ratio. This result derives from the assumptions made

    about the distribution of ice deposited on the crystal

    surface and can be partially circumvented in traditional

    mass–size methods by parameterizing the change in as-

    pect ratio (e.g., Harrington et al. 1995; Avramov et al.

    2011). One could argue that these methods are in-

    consistent with capacitance theory itself. The method of

    CL94 addresses this shortcoming of the capacitance

    model by modifying the vapor fluxes along the a and c

    axes, producing a parameterization that is more consis-

    tent with the theory and rooted in laboratory measure-

    ments. This modification allows for aspect ratio

    evolution through a second predictive equation based

    on a mass distribution hypothesis relating the growth of

    the a and c axes:

    dc

    da5 d(T)f . (2)

    The inherent growth ratio [d(T)5 ac/aa] is a ratio of thedeposition coefficients, or growth efficiencies, for the c

    and a axes. Since the deposition coefficients can be de-

    termined from laboratory measurements (e.g., Lamb

    and Scott 1972; Nelson and Knight 1998; Libbrecht

    2003), the method can accommodate new data. Because

    Eqs. (1) and (2) together are fundamentally differ-

    ent from the capacitance model, SH11 reterms it the

    Fickian-distribution method so that a clear distinction is

    drawn between the two methods of growth. Values of d

    used here are compiled from two sources: the original

    work of CL94 using laboratory data from Lamb and

    Scott (1972) that derive from crystals grown in pure

    vapor on a solid substrate, and values derived byHashino

    and Tripoli (2008), which were altered for numerical

    accuracy in the solutions and to account for the possi-

    bility of plates as the primary habit at temperatures

    below 2228C (see Bailey and Hallett 2009). The mod-ified values of d in Hashino and Tripoli (2008), larger

    (smaller) d than CL94 for plates (columns), are useful

    because they are within the limits of the measured d

    and together with best-fit values from CL94 provide

    realistic bounds on possible values of d. It is critical to

    note that neither the capacitance model nor the Fickian-

    distribution method has been well tested for situations

    in which the temperature varies or for low ice super-

    saturations (see Nelson 1994; Sheridan et al. 2009;

    SH11). Though the method has been extended to more

    complex crystal types such as rosettes and polycrystals

    (e.g., Hashino and Tripoli 2008), it is unclear how ac-

    curate the method is for these types of particles since

    comparisons to laboratory data are needed.

    The nonlinear impact of crystal shape on growth is

    embodied by an evolving capacitance in the Fickian-

    distribution method (C; Fig. 1): For the same volume

    particle, C increases nonlinearly as the aspect ratio de-

    viates from unity. This is due to the tightening of the

    vapor density gradients near crystal edges (Libbrecht

    2005, their Fig. 5, and see SH11 for a full explanation).

    While the capacitance contains this characteristic feature

    of the vapor field (Fig. 2), the aspect ratio is constant in

    the traditional capacitance model. The consequences of

    keeping aspect ratio constant are illustrated in Fig. 1: A

    single crystal is grown for 5 min at liquid saturation and at

    T5268C either with a constant or evolving aspect ratio.When f is constant, the nonlinear feedbacks in growth

    are not captured, and so the crystal grows to a smaller

    final capacitance, and therefore mass, than when f

    evolves freely. Consequently, predicting crystal aspect

    ratio is critical for capturing nonlinear changes in vapor

    diffusion.

    The comparisons of CL94 and SH11 with wind tunnel

    data suggest that a variety of crystal shapes can be used

    in the Fickian-distribution method as long as d andC are

    FIG. 1. Capacitance as a function of aspect ratio for oblate (plate)

    and prolate (column) crystals with volume equivalent to that of

    a sphere (req) of radius 50 (solid), 25 (dashed), and 10 (dotted–

    dashed) mm. Evolution of an ice particle at T 5 268C, liquid sat-uration, an initial equivalent volume spherical radius of 10 mm, and

    initial f 5 10 for the capacitance model with f constant (solidblack arrow) and Fickian distribution with evolving aspect ratio

    (solid gray arrow).

    FEBRUARY 2013 HARR INGTON ET AL . 351

  • known. Using a single crystal type (e.g., hexagon or

    dendrite) restricts the generality that is desired for a nu-

    merical cloud model. Therefore, spheroids are used as

    a first-order approximation of primary habit (f), allowing

    for the prediction of two axes. Relating changes in mass

    to changes in axis lengths requires the spheroidal volume:

    V(a, c)54

    3pa2c5

    4

    3pa3f and m(a, c)5V(a, c)ri ,

    (3)

    where V is the volume, and ri is the effective particle

    density that is less than bulk ice (rbi5 920 kg m23). The

    effective density is a parameterization of the secondary

    habits (i.e., dendrites and needles) and varies in time

    since the density deposited during growth, the depo-

    sition density [rdep; see CL94, their Eq. (42)], varies with

    temperature. Over short growth times rdep is approxi-

    mately constant (CL94) so that

    dm

    dt5 rdep

    dV

    dt. (4)

    SH11 point out that an analogous density is needed for

    sublimation [rsub; their Eq. (6)], though here rdep is used

    to signify either density. Solving Eq. (1) using an

    equivalent volume sphere requires a relation between

    changes in r and a. Equating the spheroidal and spher-

    ical volumes, then logarithmically differentiating and

    using Eq. (2) results in

    dr

    da5

    21 d

    3

    r

    a. (5)

    Thus, changes in r can be used to find changes in a and

    then changes in c [Eq. (2)].

    The equations above constitute the core of adaptive

    habit evolution using the Fickian-distribution method.

    The equations are modified for ventilation effects that

    require computation of crystal fall speeds. The method

    of SH11 is followed, and the main equations are dis-

    cussed in appendix A. In Part II, laboratory data are

    used to critique the above equations and prior mass–size

    methods. It will be shown that the above method is ro-

    bust and relatively accurate, whereas most mass–size

    relations produce crystal masses, sizes, and fall speeds

    that are generally in error.

    3. Bulk adaptive habit parameterization

    Prediction of bulk ice habit requires the particle size

    distribution-integrated forms of the mass growth equa-

    tion [Eqs. (1) and (4)] and the mass distribution hypoth-

    esis [Eq. (2)]. In forming these equations, a quantitative

    link between the distributions of the a and c axes is

    needed. Two-dimensional size spectra have been used

    for this purpose (e.g., Chen and Lamb 1999) but are

    computationally costly. Here, d is used to relate the

    a- and c-axis distributions, so that a single distribution

    can be used, making the method numerically efficient.

    The development presented here assumes the evolu-

    tion of a single habit type, and so mixes of particle

    habits (plates and needles) and complex crystal types

    (e.g., rosettes and polycrystals) are not considered, but

    could be added by allowing for at least two particle

    classes and following a method similar to Hashino and

    FIG. 2. Graphical illustration of vapor flux enhancement due to nonsphericity in the capacitance model. Each particle has the same

    volume, with vapor density contours of the same values surrounding the particle and computed for an oblate spheroid. Spherical particles

    have isometric vapor fields, and hence the vapor flux is uniform over the surface. However, as the aspect ratio f varies away from one,

    decreasing in this case, the vapor flux increases over the axis of greatest curvature (end of the a axis) and decreases over the other axis

    (c axis). This process leads to an overall increase in the diffusive flux toward the particle.

    352 JOURNAL OF THE ATMOSPHER IC SC IENCES VOLUME 70

  • Tripoli (2008). Frequently used coefficients from the

    derivations below are listed in Table 1.

    a. Axis length history

    The key to predicting the first-order effects of habit

    evolution lies in historically relating the a- and c-axis

    lengths. This is necessary because the inherent growth

    ratio d in Eq. (2) varies with temperature, and therefore

    the way a and c are related will vary in time. This vari-

    ability is accounted for by integrating Eq. (2) in time

    from a0 to a(t) and c0 to c(t), giving

    c(t)5a*ad*(t) and d*5

    ða05a(t)a05a

    0

    d(t) d ln(a0)

    ða05a(t)a05a

    0

    d ln(a0), (6)

    where a*5 c0a2d*0 , and d* is a time average of the in-

    herent growth ratio over the particle growth history, also

    described in CL94. Note that the weighting is loga-

    rithmic, which is a consequence of the relationship be-

    tween dc/c and da/a in Eq. (2). The initial aspect ratio

    f0 and size a0 of the ice particles must be chosen. Initial

    aspect ratios that are not unity can be accommodated by

    setting c0 5 f0a0 in a*. Although some recent labora-tory data suggest that initially frozen drops at T 52408C have aspect ratios of around 1.1 (López andÁvila 2012), given that data on the initial ice aspect ratio

    is sparse and that laboratory data suggest aspect ratios

    near one, it is assumed here that initially formed crystals

    are isometric, and so a*5 a12d*0 .

    Because particles are initially assumed to be iso-

    metric, d* is initially equal to one. However, the in-

    herent growth ratio d is typically not unity and varies

    with temperature, and so d* will also vary in time. This is

    an important departure from traditional bulk methods

    that implicitly fix d* at a particular value for a given

    habit type. In the method proposed here, the relation

    between c and a will also evolve. Since d* is a weighted

    time average of d, it is naturally constrained to lie within

    the limits of d derived from laboratory data (about 0.27

    and 2.1). Note the distinction between d and d*: d is

    a measure of the mass distributed over the a and c axes

    during growth (one time step), whereas d* relates the two

    axis lengths over time and allows for the use of a single

    size spectrum in either the c- or a- axis length. It is worth

    noting that d* is not a prognostic variable, but is diagnosed

    from separate prediction of the c- and a-axis lengths. The

    parameterization depends on the exactness of this his-

    torical relation (d*), which is demonstrated in Part II.

    Using Eqs. (3) and (6), the ice crystal mass can be

    written in the traditional mass–size form

    m(a)5V(a)ri 5amabm , (7)

    where the coefficients am 5 (4/3)pa*ri and bm 5 21 d*are for a platelike crystal. This result shows that the

    mass–size relations are directly connected to the his-

    torical evolution of the particle through d*. Conse-

    quently, empirical mass–size relations should always be

    used with caution as the entire integrated history of the

    particle’s evolution is contained in the leading co-

    efficient and the exponent. Because equivalent volume

    spheres are used to solve for the mass changes over time

    (section 2), a historical relation between r and a is

    needed and can be found by equating spheroidal and

    spherical volumes and then using Eq. (6):

    a5arrbr , (8)

    where ar 5a21/bm*

    and br 5 3/bm.

    b. Axis distributions

    The bulk habit growth equations require a priori

    forms of the particle distributions. Since c and a are re-

    lated to each other through Eq. (6), the distribution of

    only one axis length is needed, and the a-axis length is

    used here. Following other bulk modeling approaches

    (e.g., Walko et al. 1995; Thompson et al. 2004), the

    concentration of crystals conforms to a modified gamma

    distribution of a:

    n(a)5NiG(n)

    �a

    an

    �n21 1an

    exp

    �2

    a

    an

    �, (9)

    where Ni is the number density of ice crystals, G(n) is thegamma function, n is the distribution ‘‘shape’’ parameter,

    and an is the characteristic a-axis length that is related to

    TABLE 1. Coefficients resulting from derivations presented in the

    main text and appendix.

    Bulk model coefficients

    Variable Definition Equation number

    a* a12d*0 (6)

    aa a1/3* (B19)

    acap 1 (platelike) or a* (columnlike) (B13)

    am43pa*ri (7)

    ar a21/bm*

    (8)

    ay43pa* (B33)

    ba21 d3

    (5)

    ba* (2 1 d*)/3 (B19)bm 2 1 d* (7)br 3/bm (8)

    dcap 1 (platelike) or d* (columnlike) (B13)

    FEBRUARY 2013 HARR INGTON ET AL . 353

  • the mean a-axis length, and the a-axis mixing ratio (see

    section 3e). It is worth noting that generalized distri-

    butions like that in Thompson et al. (2008) could also

    be used.

    To define the aspect ratio requires the c-axis distribu-

    tion while the equivalent volume sphere distribution (r) is

    needed for vapor growth calculations (see section 2).

    Therefore, it is necessary to change variables from a to

    that of r or c in Eq. (9), which can be done analytically by

    using d*. The resulting distributions are similar toEq. (9).

    The general form of the distribution conversion and the

    distributions of the c and r axes [Eqs. (B6) and (B5), re-

    spectively] are given in section a of appendix B. Each has

    characteristic lengths (cn and rn) related to the mean axis

    length that characterize the mean particle shape.

    Eulerian models require conserved variables, and so

    four quantities are predicted in the bulk adaptive habit

    method: The number mixing ratio (Ni/ra), the ice mass

    mixing ratio (qi), and the mixing ratios of the a and c

    axes, defined respectively as

    A5nNira

    an,

    C5nNira

    cn , (10)

    where ra is the air density. These equations follow

    from the definition of the mean axis length [a5(1/Ni)

    Ð ‘0 an(a) da5 an[G(n1 1)]/[G(n)]5 nan, Eq. (B4)].

    The axis length mixing ratios are a measure of the

    summed lengths of a and c for all the ice particles per kg

    of air. It is beyond the scope of the present work to

    demonstrate that the above quantities are conserved in

    an Eulerian framework, since testing is carried out

    using a parcel model in Part II; however, a short dis-

    cussion of this issue is provided in section 3e.

    c. Mass mixing ratio growth and mass distributionrelationship for ice particle distributions

    Knowing the particle distributions [Eqs. (9), (B5), and

    (B6)] and the historical relation between the axis lengths

    [Eq. (6)] makes it possible to derive equations for the

    evolution of the mass mixing ratio and the axis length

    mixing ratios. Because the axis mixing ratios are pro-

    portional to an and cn, the characteristic sizes are used

    throughout to represent mean habit evolution. Evolving

    qi follows a procedure similar to that for single particles

    (section 2a): The mass growth equation is combined

    with an equation describing the change in characteristic

    equivalent volume radius. Once the change in rn is

    known, aspect ratio evolution can then be computed

    using relationships between the different axis lengths.

    The evolution of qi by vapor diffusion is given by

    dqidt

    51

    ra

    ð‘0

    dm

    dtn(a) da . (11)

    To integrate Eq. (1) over n(a) [Eq. (9)] the capacitance

    (C) must be written in terms of the a-axis length alone.

    This relationship is derived by using Eq. (6) and fitting

    the shape factor with a power law in a, as is done in

    section b of appendix B. Integrating over the size

    spectrum [Eq. (11)] produces an equation in an alone

    [see section b of appendix B; Eq. (B13)]. Integrating

    the resulting expression [Eq. (B13)] in time is compli-

    cated because the mass and capacitance both depend

    on an and, implicitly cn, both of which vary over a time

    step (see section b of appendix B). As for single-particle

    mass (section 2), tests show that evolving the charac-

    teristic equivalent volume spherical radius (rn) rather

    than an or cn to determine the mass mixing ratio change

    over a time step provides the most accuracy in com-

    parison to numerical solutions. The differential equation

    for qi in terms of rn is shown in section d of appendix B

    to be

    dqidt

    5Nira

    4prnfs(rn)Gisu,i , (12)

    where fs(rn)5C(an)/rn is a representative shape factor.To make this equation amenable to integration, qi must

    be related to rn, which can be done by integrating the

    particle mass over n(r) [Eq. (B5)]:

    qi 54

    3pr3nriNi

    G(n1 3/br)

    G(n)5 riVt , (13)

    with ri the volume-weighted mean effective density (or

    ‘‘mean effective density’’; see section d of appendix B)

    and Vt the total particle volume. Using this relationship,

    changes inmass mixing ratio can be related to changes in

    Vt and rn in a similar fashion as for single particles [Eq.

    (4)] through

    dqidt

    5 rdepdVtdt

    5 4prdepNiG[n1 (3/br)]

    G(n)r2ndrndt

    . (14)

    Equation (14) along with Eq. (12) can now be solved to

    advance the characteristic equivalent radius rn, and

    hence the mass mixing ratio, in time (see section d of

    appendix B).

    Evolving the characteristic axis lengths requires

    a mass distribution relationship connecting changes in rnto changes in cn and an. Such a relation can be derived

    using Eq. (5) (see section c of appendix B):

    354 JOURNAL OF THE ATMOSPHER IC SC IENCES VOLUME 70

  • drndan

    5

    �21 d

    3

    �rnan

    . (15)

    The mass distribution relationship connecting changes

    in an to cn is shown in section c of appendix B to be

    dcndan

    5 dcnan

    (16)

    and closes the equation set. The procedure used to solve

    these equations for mean habit evolution is provided in

    section 3e.

    d. Fall speed and ventilation

    The bulk habit method outlined above includes nei-

    ther ventilation effects nor estimates of particle fall

    speeds, both of which are needed for any bulk parame-

    terization. Computing the particle fall speeds requires

    choosing appropriate coefficients for the Reynolds

    number (see section e of appendix B), which thenmakes

    it possible to find the bulk fall speed by combining Eqs.

    (A1) and (A3) and integrating over the size spectrum:

    yt 51

    Ni

    ðNReharaL

    n(a) da5NRe(an)haraL(an)

    G(n1 bnbm 2 bl)

    G(n),

    (17)

    where the coefficients are defined in appendix A.

    Eulerian models also require a mass- (ytm) and length-

    weighted (ytl) fall speed, and these are given in section e

    of appendix B.

    Computing the bulk ventilation coefficients requires

    diagnosing the power-law coefficients (gy and gt) that

    appear in the equation for ventilation [Eq. (A4)], and

    the process is explained in section e of appendix B.

    These coefficients multiply the vapor diffusivity (f yDy)

    and thermal conductivity (f tKt) that appear in Gi [Eq.

    (11)] and are determined by appropriately integrating

    over the a-axis distribution (section e of appendix B):

    f y 5 by11 by2 N1/3Sc NRe(an)

    1/2h ig

    yf cap,y ,

    f t 5 bt11 bt2 N1/3Pr NRe(an)

    1/2h ig

    yf cap,t , (18)

    with all coefficients defined by Eq. (B42) in section e of

    appendix B and in appendix A.

    Ventilation affects each axis length differently, and

    so bulk versions of the axis-dependent ventilation co-

    efficients [i.e., Eq. (A5)] are also needed. Using a pro-

    cedure similar to that above (see section e of appendix B)

    leads to the following coefficients for the a and c axes,

    respectively:

    fya5by11by2 N1/3Sc NRe(an)

    1/2h ig

    ya21/3* a

    1/3(12d*)n

    h igy/2f cap,a,

    f yc5by11by2 N1/3Sc NRe(an)

    1/2h ig

    ya2/3* a

    2/3(d*21)n

    h igy/2f cap,c ,

    (19)

    and the terms f cap,a and f cap,c are defined by Eq. (B46) in

    section e of appendix B.

    The effects of ventilation on the evolution of themean

    axis lengths are then estimated following the same

    procedure that is used for single-particle axis lengths

    [Eq. (A6)]:

    dyb5 df ycf ya

    . (20)

    The value of dyb is the bulk value of the inherent growth

    ratio modified for ventilation effects. Including axis-

    dependent ventilation is now accomplished by replacing

    dwith dyb in all of the bulk equations derived above. It is

    critical to keep in mind that only d is modified whereas

    d* remains as defined in Eq. (6).

    e. Procedure for evolving habits

    Predicting changes in mass mixing ratio, axis lengths,

    fall speeds, and density can now be accomplished using

    the equations above when solved over a time step, the

    details of which are provided in the appendices. The

    derivation of the method is complicated, so an overview

    of the equations used to evolve bulk habits is warranted.

    Moreover, the equations are cast in a form suitable for

    use in Eulerian models since the method is designed

    ultimately for this purpose. Themain equations used are

    also summarized in Tables 2 and 3.

    Eulerian models use conserved quantities, and the

    mass, number, and axis length mixing ratios (qi,Ni/ra,A,

    and C, respectively) are used here. To compute the

    evolution of mass mixing ratio and axis lengths, a num-

    ber of quantities are needed at the beginning of the time

    step t and can be extracted from qi, A, and C. Table 2

    provides a summary of the main initial variables. At the

    beginning of a time step, cn and an can be determined

    from the prognosed axis length mixing ratios using Eq.

    (10). With these, d*(t) can be diagnosed by taking the

    natural logarithm of cn(t)5a*an(t)d*(t)

    [Eq. (B6)] and

    replacing a*5 a(12d*)

    0 :

    d*(t)5ln[cn(t)]2 ln(a0)

    ln[an(t)]2 ln(a0). (21)

    Once d* is known at the beginning of a time step, the

    characteristic equivalent volume spherical radius rn(t)

    can be found from Eq. (B5), and the mean ice particle

    FEBRUARY 2013 HARR INGTON ET AL . 355

  • density ri(t) can be solved for by using the ice mass

    mixing ratio qi(t) and inverting Eq. (13). To compute

    vapor diffusion, quantities that depend on the charac-

    teristic lengths are needed. The average capacitance

    C(an) and the average shape factor f s(an) are de-

    termined using Eqs. (B13) and (B14) (see Table 2). The

    vapor and thermal diffusivities that appear in the growth

    equation [Gi in Eq. (12)] are modified for ventilation

    effects using Eq. (18) (Table 2).

    Time stepping of the variables can now be done, and

    the main equations are summarized in Table 3. For

    reasons given in section 3c and in SH11, particle size and

    TABLE 2. Summary of variables: Beginning of time step. Only the mixing ratios (Ni/ra, qi,A, and C) are stored externally to the growth

    calculations. To compute the change in qi, A, and C over a time step, initial values of various quantities must be derived from these

    quantities. The initial variables at the beginning of a time step used in the growth model are given here, and a description in section 3e.

    Equation Purpose Equation originates from

    an(t)5ranNi

    A(t) Characteristic a-axis length Eq. (10)

    cn(t)5ranNi

    C(t) Characteristic c-axis length Eq. (10)

    d*(t)5ln[cn(t)]2 ln(a0)

    ln[an(t)]2 ln(a0)Initial value of d*, needed for rn and growth calculations Eq. (21)

    rn(t)5a21/brr a

    1/brn Initial equivalent spherical characteristic radius Eq. (B5)

    ri 53qi(t)

    4prn3(t)Ni

    G(n)

    G(n1 3/br)Initial average density Eq. (13)

    fs(rn)5C(an)

    rn5

    1

    rnacapa

    dcapn fs(an) Shape factor for vapor growth Eq. (B14)

    Dyf y 5Dyfby1 1by2[N1/3Sc NRe(an)1/2]gy f cap,ygKTf t 5KTfbt1 1bt2[N1/3Pr NRe(an)1/2]gy f cap,tg

    Ventilation effects on diffusion, used in Gi Eq. (18)

    dyb 5 df yc

    f yaVentilation modification of d Eqs. (19) and (20)

    TABLE 3. Summary of prognostic variables: End of time step. The growthmethod evolves the icemassmixing ratio (qi), and the a- and c-

    axis mixing ratios (A andC) over a time step. This table contains themain equations in the order that they are used. Note that initial values

    of the variables from Table 2 are assumed. At the end of the calculations, only the three mixing ratios are passed back and stored.

    Equation Purpose

    Equation originates

    from

    rn(t1Dt)5

    "r2n(t) 1

    2fs(rn)Gisu,irdep

    G(n)

    G(n13/br)Dt

    #1/2 Needed to find qi and axislengths at end of Dt

    Eq. (B23)

    an(t1Dt)5 an(t)

    �rn(t 1 Dt)

    rn(t)

    �3/(21dyb)Characteristic a-axis length

    at end of DtEq. (B26)

    d*(t1Dt)53 ln[rn(t1Dt)]2 2 ln[an(t1Dt)]2 ln(a0)

    ln[an(t1Dt)]2 ln(a0)Needed to find cn at end of Dt Eq. (B27)

    cn(t1Dt)5a*(t1Dt)ad*(t1Dt)n cn calculation when crystals are

    columnar (d* . 1)Eq. (B6)

    f(t1Dt)

    f(t)’�r3n(t1Dt)

    r3n(t)

    �(dyb21)/(dyb12)cn(t1Dt)

    5f(t1Dt)an(t1Dt)G(n)

    G(n1 d*2 1)

    cn calculation when platelike

    (d* , 1): first find f at end ofDt, next find cn at end of Dt

    Eqs. (B31) and (B32)

    ri(t1Dt)5 (12wy)rdep 1wyri(t)ri Change in mean bulk ice density.Used to find new value of ice

    mixing ratio

    Eq. (B25)

    qi(t1Dt)5 qi(t)1Dqi 5qi(t)1 ri(t1Dt)Vt(t1Dt)2 ri(t)Vt(t) Ice mass mixing ratio at theend of Dt

    Eq. (B24)

    A(t1Dt)5 (nNi/ra)an(t1Dt)C(t1Dt)5 (nNi/ra)cn(t1Dt)

    a- and c-axis mixing ratio at

    the end of DtEq. (10)

    356 JOURNAL OF THE ATMOSPHER IC SC IENCES VOLUME 70

  • mass changes are computed by finding the change in rnusing Eq. (B23) giving rn(t1 Dt). Knowing rn at the endof a time step allows an(t 1 Dt) to be determined usingEq. (B26). Including axis ratio-dependent ventilation is

    then accomplished by replacing d in Eq. (B26) with

    dyb [Eq. (20)]. The change in cn can also be computed,

    but to do so requires knowing d*(t 1 Dt), which is di-agnosed from Eq. (B27) since rn(t 1 Dt) and an(t 1 Dt)are known. For columnlike crystals, cn(t 1 Dt) is de-termined from the relation between cn and an in Eq.

    (B6). For platelike crystals, the change in mean aspect

    ratio is first computed, and then cn(t1 Dt) is found fromEq. (B32). At this point, the axis mixing ratios are re-

    computed from Eq. (10).

    Finally, the mass mixing ratio at the end of a time step

    is computed. To do so requires knowing how the mean

    ice density has changed over time. The new density is

    a volume-weighted average between the density at the

    prior time step and the density added during growth

    given by Eq. (B25). Once ri(t1Dt) is known, it is thenpossible to compute the change in ice mass mixing ratio

    over a time step through Eq. (B24). The mixing ratios qi,

    A, and C are all now known at the end of a time step.

    In the parcel model tests described in Part II of this

    paper, our goal is to illustrate that the average effects of

    adaptive habit evolution can be captured with the pro-

    cedure described above. Other issues emerge when the

    method is implemented in an Eulerian framework. For

    example, separate advection of the a- and c-axis length

    mixing ratios could potentially lead to errors in the

    evolution of axis ratio. However, tests indicate that such

    errors are small; this is a subject of a forthcoming paper

    on the Eulerian implementation of the method. Other

    issues also arise: Ice can be nucleated at multiple loca-

    tions within the cloud, and regions of differing ice classes

    can be mixed together. Furthermore, the adaptive habit

    method needs to be connected in a physically accurate

    way to both riming and aggregation processes. These

    issues are beyond the scope of the present work, but are

    discussed briefly in Part II and in more depth in our

    forthcoming paper on the Eulerian implementation and

    testing.

    4. Summary, remarks, and future development

    Prior methods of parameterizing ice mass growth

    from vapor have used the traditional capacitance

    model; however, the capacitance depends on the major

    semiaxis length (either a or c) and the aspect ratio

    (f5 c/a) of the particles. Most bulk cloud models closethe equations by relating mass and axis length through

    a mass–size relation derived from in situ data while

    keeping the aspect ratio constant in the capacitance.

    A constant f severs the nonlinear link between aspect

    ratio changes and increases in vapor growth leading to

    compounding errors in ice mass evolution. Because

    mass–size relations are derived for particular ice habits,

    cloud models then account for habit changes through

    the use of multiple ice classes (e.g., Walko et al. 1995;

    Woods et al. 2007) or assume a constant particle habit

    (e.g., Reisner et al. 1998; Fridlind et al. 2007; Thompson

    et al. 2004; Avramov and Harrington 2010). A more

    recent trend in cloud modeling is to predict a number

    of particle properties instead of different ice classes.

    Doing so produces a more natural evolution of the

    particle’s properties over time and avoids transfer

    functions between ice classes which are poorly con-

    strained and often arbitrary. Along these lines, a bulk

    method for parameterizing ice habit evolution with

    four prognostic variables was developed that captures

    the nonlinearity in ice vapor growth through aspect

    ratio changes following SH11. The habit method could

    also be used for models that attempt to predict distri-

    bution shape (Milbrandt andYau 2005) or for Eulerian

    bin microphysical codes (e.g., Reisin et al. 1996). In

    Part II of this work, the accuracy of the method is ex-

    amined, and the limitations of the method are com-

    mented upon.

    Acknowledgments. We are grateful for the detailed

    comments provided by three anonymous reviewers,

    which substantially improved this manuscript. K. Sulia

    and J. Harrington would like to thank the National

    Science Foundation for support under Grants ATM-

    0639542 and AGS-0951807. In addition, J. Harrington

    was supported in part by the Department of Energy

    under Grant DE-FG02-05ER64058. K. Sulia was sup-

    ported in part by an award from the Department of

    Energy (DOE) Office of Science Graduate Fellowship

    Program (DOE SCGF). The DOE SCGF Program was

    made possible in part by the American Recovery and

    Reinvestment Act of 2009. The DOE SCGF program

    is administered by the Oak Ridge Institute for Science

    and Education for the DOE. ORISE is managed by

    Oak Ridge Associated Universities (ORAU) under

    DOE Contract Number DE-AC05-06OR23100. All

    opinions expressed in this paper are the author’s and

    do not necessarily reflect the policies and views of

    DOE, ORAU, or ORISE. H. Morrison was partially

    supported by the NOAA Grant NA08OAR4310543;

    U.S. DOE ARM DE-FG02-08ER64574; U.S. DOE

    ASR DE-SC0005336, sub-awarded through NASA

    NNX12AH90G; and the NSF Science and Technology

    Center for Multiscale Modeling of Atmospheric Pro-

    cesses (CMMAP), managed by Colorado State Univer-

    sity under Cooperative Agreement ATM-0425247.

    FEBRUARY 2013 HARR INGTON ET AL . 357

  • APPENDIX A

    Fall Speed and Ventilation Effects

    A unique feature of the CL94 method is that venti-

    lation during particle sedimentation affects each axis

    differently, and this reproduces the observed tendency

    of the crystal minor axis to asymptote to a particular size

    (see their Fig. 6). Calculating ventilation effects requires

    knowledge of nonspherical particle fall speeds. Sulia and

    Harrington (2011) is followed where a combination of

    the methods of Bohm (1992) and Mitchell (1996) are

    used, which depend on the particle Reynolds number:

    yt 5NRehara

    1

    L, (A1)

    where ra is the air density, ha is the dynamic viscosity,

    and L is an appropriate particle length scale. Following

    Bohm (1992), L 5 2a for a plate, 2ffiffiffiffiffiffiffiffiffi(ac)

    pfor a column,

    and 2r for a sphere as they produce the best accuracy

    in comparison to wind tunnel data of growing ice

    particles (see SH11). The particle Reynolds number

    (NRe 5 amXbm) is computed from a power-law fit to theBest number X using the fit coefficients (am and bm)

    from Mitchell (1996). The Best number depends only

    on particle geometry and the form given in Bohm

    (1992) but corrected for buoyancy effects is used:

    X 5 [(2mgraL2)/(Aeh

    2a)]q

    3/4e [(ri 2 ra)/ri], where ri is the

    ice particle density, m is the crystal mass, g is the grav-

    itational acceleration, Ae is the projected area of the

    crystal [pa2(ri/rbi)2/3 for platelike and pac for columnlike],

    and qe is the ratio of the projected area of the crystal to the

    projected area of the spheroid (Ae/A; see SH11). The Best

    number is recast in a form more conducive to bulk model

    development by exposing the a-axis length dependence

    using Eq. (6) in the relations for L and Ae:

    X5 xBmL2

    Aewhere xB5

    2(ri2 ra)graq3/4e

    rih2a

    ,

    5 xnabn , (A2)

    where Ae 5 aaaba , L5 alabl , xn 5 xB(4/3)[(pria*a2l )/aa],

    bn5 d*1 21 2bl2 ba, and al5 2, bl5 1; aa5 p(ri/rbi)2/3

    for platelike crystals; and al 5 2a1/3* , bl 5 (d*1 2)/3, aa5a*p, and ba5 d*1 1 for columnar crystals. TheReynoldsnumber can now be written in terms of the a axis alone:

    NRe 5 amXbm 5 amx

    bm

    n abnbm . (A3)

    These forms of the Best and Reynolds numbers can be

    integrated over the a-axis distribution.

    Ventilation effects on mass growth can be determined

    if the vapor diffusivityDy and thermal conductivity of air

    Kt are multiplied by respective ventilation factors:

    fy 5 by11 by2Ngy/3

    Sh Ngy/2

    Re ,

    ft 5 bt11 bt2Ngt/3

    Pr Ngt/2

    Re , (A4)

    where NSh 5 ha/(raDy) is the Sherwood number, andNPr 5 ha/(raKt) is the Prandtl number. The coefficientsabove are given in CL94.

    Axis ventilation follows CL94. Nonspherical particles

    tend to fall with their major axis perpendicular to the fall

    direction, and the crystal edges experience a stronger

    ventilation effect (see SH11), which is approximated for

    a and c, respectively, as

    fya5 by11 by2Ngy/3

    Sh Ngy/2

    Re

    �ar

    �gy/2,

    fyc5 by11 by2Ngy/3

    Sh Ngy/2

    Re

    �cr

    �gy/2, (A5)

    where r is the radius of an equivalent volume sphere.

    Chen and Lamb (1994) shows that axis ventilation

    modifies the inherent growth ratio in the following

    manner:

    dy(T)5 d(T)fycfya

    , (A6)

    where d is replaced by dy in Eq. (2). While these re-

    lationships are hypothesized, they produce relatively

    good agreement between wind tunnel measurements

    and modeled evolution of crystal mass, axis lengths, and

    fall speed (SH11). Note that thermal effects are not in-

    cluded, because dy is associated with the vapor re-

    distribution and not the thermal diffusion process

    (CL94).

    APPENDIX B

    Bulk Parameterization Details

    a. Distribution relations

    The p moment of the a-axis distribution [Eq. (9)] is

    given by (see Walko et al. 1995)

    I(p)5

    ð‘0apn(a) da5Nia

    pn

    G(n1 p)

    G(n). (B1)

    Since the habit method uses both the c axis and the

    equivalent volume spherical radius r, it is necessary to

    find the distributions andmoments of c and r. The key to

    358 JOURNAL OF THE ATMOSPHER IC SC IENCES VOLUME 70

  • the bulk habit method is the historical aspect ratio

    contained in Eqs. (6) and (8), which connects the vari-

    ation in a to the other axis lengths. The relationship

    between a and either r or c has the general form

    a5addbd , (B2)

    where ad and bd are appropriate coefficients defined

    below, and d is either r or c. Changing variables from a to

    d in Eq. (9) gives

    n(d)5NiG(n)

    bd

    �d

    dn

    �nbd21 1

    dnexp

    �2

    �d

    dn

    �bd�,

    an5addbd

    n . (B3)

    Note the relatively simple relationship between an and

    dn. The moments of the converted size distribution are

    Id(p)5

    ð‘0dpn(d) d(d)5Nid

    pn

    G(n1 p/bd)

    G(n). (B4)

    The distribution conversion and moments can now be

    used to convert the a-axis distribution to a distribution of

    other length scales. Equations (8) and (B3) produce the

    distribution of r:

    n(r)5NiG(n)

    br

    �r

    rn

    �nbr21 1

    rnexp

    �2

    �r

    rn

    �br�,

    an5arrbr

    n . (B5)

    Converting from the a- to the c-axis distribution can be

    done in a similar fashion using Eqs. (6) and (B3):

    n(c)5NiG(n)

    1

    d*

    �c

    cn

    �(n/d*)21 1cn

    exp

    �2

    �c

    cn

    �1/d*�,

    cn 5a*ad*n . (B6)

    b. Integration of mass growth equation

    To integrate Eq. (11) analytically, the capacitance

    must be written in terms of a alone. Using Eq. (6), the

    capacitance [Eq. (1)] can be written in a general form for

    either columnlike or platelike particles:

    C(a,f)5acapadcap fs(f) , (B7)

    where fs is the shape factor, acap 5 1, dcap 5 1 forplatelike crystals, and acap 5 a*, dcap 5 d* for columnarcrystals. The aspect ratio can be written in terms of

    a alone using the time–historical relationship between

    a and c [Eq. (6)]:

    f(a)5a*ad*21 . (B8)

    The mathematical forms of the shape factors for

    spheroids are complex enough that analytical in-

    tegration over the a-axis distribution is not possible. A

    simple form for fs is sought, and so a power-law fit is

    used:

    fs(f)5 a1fb1 1 a2f

    b2 , (B9)

    where a1, a2, b1, and b2 are fit coefficients with values

    given inTableB1. FigureB1 shows that Eq. (B9) provides

    an accurate estimate of the shape factor over a broad

    range of aspect ratios. Combining Eqs. (B7), (B8), and

    (B9) gives the capacitance in terms of a only:

    C(a)5 c1ad1 1 c2a

    d2 , (B10)

    with the coefficients

    c15 a1acapab1

    *c25 a2acapa

    b2

    *d1 5b1(d*2 1)1 dcap d25 b2(d*2 1)1 dcap .

    (B11)

    With these relations, the growth equation [Eq. (11)] can

    now be integrated:

    dqidt

    51

    ra

    ð‘04pC(a)Gisu,in(a) da ,

    51

    raNi4pGisu,i

    �c1a

    d1

    nG(n1 d1)

    G(n)1 c2a

    d2

    nG(n1d2)

    G(n)

    �,

    (B12)

    where Eq. (B1) is used. It is convenient to cast the

    growth equation given by Eq. (B12) in a form similar to

    that of a single particle where the capacitance is written

    as the maximum semidimension (either c or a) multi-

    plied by a shape factor fs. Equations (6) and (B7) suggest

    the following form:

    dqidt

    5Nira

    4pacapadcapn fs(an)Gisu,i , (B13)

    TABLE B1. Power-law fit coefficients for the shape factor [fs; Eq.

    (B9)] as a function of aspect ratio.

    Fit coefficients

    a1 a2 b1 b2

    Oblate 0.636 942 7 0.363 057 0.0 0.95

    Prolate 0.571 428 5 0.428 571 21.0 20.18

    FEBRUARY 2013 HARR INGTON ET AL . 359

  • with the distribution-averaged capacitance as

    C5acapadcapn fs(an) , (B14)

    where

    fs(an)5c1acap

    ad12d

    capn

    G(n1 d1)

    G(n)1

    c2acap

    ad22d

    capn

    G(n1 d2)

    G(n)

    5G(n1 1)

    G(n)

    ð‘0fs(a)an(a) dað‘0an(a) da

    (B15)

    and is similar to the a-axis-weighted average shape fac-

    tor for either plates or columns as the rightmost equa-

    tion indicates.

    c. Mass distribution relations for particle distributions

    Since bulk microphysical models only predict or di-

    agnose quantities characteristic of the entire particle

    distribution, a way to relate changes in the characteristic

    a, c, and r axes to one another is needed if a measure of

    particle aspect ratio is to be predicted. Using Eq. (2) and

    the growth history [Eq. (6)] a historical relation between

    changes in c and a can be found:

    dc

    dt5 da*a

    d*21da

    dt. (B16)

    Integrating over the a-axis distribution and realizing that

    n(a)da 5 n(c)dc gives

    dcndt

    5 da*ad*n

    1

    an

    dandt

    , (B17)

    which can be simplified by using the relationship be-

    tween cn and an [Eq. (B6)]:

    dcndan

    5 dcnan

    . (B18)

    This is given as Eq. (16) in the main text.

    Relating rn and an is required for vapor growth cal-

    culations. The relation can be derived by first rewriting

    Eq. (5) using Eq. (8):

    dr

    dt5baaaa

    ba*21da

    dt, (B19)

    where aa 5a1/3* , ba 5 (2 1 d)/3, and ba* 5 (21 d*)/3.Integrating this equation over the a-axis distribution and

    realizing that n(a)da 5 n(c)dr,

    drndt

    5baaaaba*

    n1

    an

    dandt

    . (B20)

    Inverting Eq. (B5) gives rn 5aaaba*n , and so the above

    equation becomes

    drndan

    5

    �21 d

    3

    �rnan

    . (B21)

    This is given as Eq. (15) in the main text.

    d. Mass mixing ratio, size, and aspect ratio evolution

    To evolvemassmixing ratio, size, and aspect ratio, Eq.

    (B13) must be integrated in time along with the equa-

    tions from the subsection above. This is done analyti-

    cally here, but since both the a and c axes vary, a solution

    to the vapor diffusion equation is problematic. As dis-

    cussed in sections 2 and 3, this problem is avoided by

    following SH11 and solving for the change in the char-

    acteristic equivalent volume radius rn. The vapor diffu-

    sion equation for mass mixing ratio [Eq. (B13)] is first

    recast into an equation for the distribution of equivalent

    volume spheres:

    dqidt

    5Nira

    4prnfs(rn)Gisu,i , (B22)

    where fs(rn)5 [C(an)/rn] is a representative shape fac-tor, and this is Eq. (12) in the main text. Equation (B22)

    FIG. B1. Shape factor as a function of aspect ratio. The exact

    expression for oblate and prolate spheroids along with the poly-

    nomial fit are shown.

    360 JOURNAL OF THE ATMOSPHER IC SC IENCES VOLUME 70

  • is set equal to Eq. (14) and integrated keeping temper-

    ature, pressure, equivalent spherical shape factor fs(rn),

    and ice supersaturation constant over a time step:

    rn(t1Dt)5

    "r2n(t)1

    2fs(rn)Gisu,irdep

    G(n)

    G(n1 3/br)Dt

    #1/2.

    (B23)

    It is worth noting that analytical integration is not re-

    quired (numerical methods can be employed; e.g.,

    Walko et al. 2000), and supersaturation does not need to

    remain constant; a simple method is chosen here to il-

    lustrate the robust nature of the parameterization.

    Calculating the change in mass mixing ratio is straight-

    forward since the change in rn is known. Integrating

    Eq. (14) over a time step yields

    Dqi5 rdepDVt 5 rdep[Vt(t1Dt)2Vt(t)]

    5 ri(t1Dt)Vt(t1Dt)2 ri(t)Vt(t) . (B24)

    Knowing the change in rn over time step allows the

    change in total volume to be computed throughEq. (14),

    and hence mass mixing ratio can be computed. The

    rightmost equation is arrived at by integrating Eq. (13),

    enabling the diagnoses of the mean effective density at

    the end of a time step:

    ri(t1Dt)5 (12wy)rdep1wyri(t) , (B25)

    where wy 5 Vt(t)/Vt(t 1 Dt) is a weighting factor.Finding the change in the characteristic a axis is also

    straightforward as changes in rn are related to changes in

    an through Eq. (15). Integrating this equation over

    a time step gives

    an(t1Dt)5 an(t)

    �rn(t1Dt)

    rn(t)

    �3/(21d). (B26)

    Note that the temporal change in an depends on two

    things: The change in the overall volume, which is pro-

    vided by the change in rn, and how that volume (or mass)

    change is distributed along the a axis, which is controlled

    by d.

    Unlike the single-particle model (CL94; SH11), cal-

    culating the change in the c-axis length is not as simple as

    using the inherent growth ratio d to estimate the change

    over time because each ice particle in the distribution

    has a different relationship between the c and a axis.

    This is why a historical relation between c and a is

    needed and provided through d*. The change in the

    characteristic c axis can be estimated as long as the

    change in d* is known, which is computed from rn and an.

    Starting with Eq. (B5) replaced with the definitions of

    ar 5a21/(21d*)

    *, a*5 a

    12d*+

    , and br 5 3/(2 1 d*), andtaking the natural logarithm, one can show that

    d*(t1Dt)53 ln[rn(t1Dt)]2 2 ln[an(t1Dt)]2 ln(a0)

    ln[an(t1Dt)]2 ln(a0),

    (B27)

    which is the value of d* at the end of the time step. It is

    now possible to use d*(t 1 Dt) to estimate cn(t 1 Dt)using Eq. (B6), which relates the characteristic c- and a-

    axis lengths.

    Estimating cn with Eq. (B6) was found to be most

    accurate for columnar crystal (d* . 1) in comparisonwith the detailed model of SH11. For platelike crystals

    (d* , 1), comparisons with SH11 showed it is more ac-curate to mimic the method used for single particles

    (CL94) and diagnose the characteristic c axis from the

    average aspect ratio:

    f51

    Ni

    ð‘0f(a)n(a) da5a*a

    d*21n

    G(n1 d*2 1)

    G(n), (B28)

    where Eq. (B8) is employed. With the relationship be-

    tween an and rn [Eq. (B5)] it is possible to show that

    f5a*ad*21r (r

    3n)

    (d21)/(d12)G(n1 d*2 1)

    G(n). (B29)

    Forming the ratio of f before and after one time step,

    f(t1Dt)

    f(t)5

    �r3n(t1Dt)

    r3n(t)

    �(d21)/(d12)

    3G[n1 d*(t1Dt)2 1]

    G[n1 d*(t)2 1]

    ad*(t1Dt)21r

    ad*(t)21r

    .

    (B30)

    The ratios of ar and the gamma functions in Eq. (B30)

    are nearly unity since d* does not change drastically

    when Dt , 20 s. With this approximation, the relation-ship between the average aspect ratio and the volume

    becomes

    f(t1Dt)

    f(t)’�r3n(t1Dt)

    r3n(t)

    �(d21)/(d12). (B31)

    Note that this relationship is nearly identical to that for

    single particles given in CL94 and in SH11. For platelike

    crystals, the characteristic c-axis length can now be re-

    lated to both themean aspect ratio and an by using the c-

    axis distribution [Eqs. (B6) and (B28)]:

    FEBRUARY 2013 HARR INGTON ET AL . 361

  • cn(t1Dt)5f(t1Dt)an(t1Dt)G(n)

    G(n1 d*2 1). (B32)

    Since f and an are known, cn can be computed after

    a time step as well.

    Diagnosing the mean effective density at the begin-

    ning of a time step requires one final relationship, and

    that is between ri and an. Integrating the spheroidal

    mass [Eq. (7)] over the a-axis distribution gives a second

    form of the mass mixing ratio:

    qi51

    ra

    ð‘0m(a)n(a) da5Niayria

    bm

    nG(n1bm)

    G(n)

    ri 5

    ð‘0ria

    bmn(a) dað‘

    0abmn(a) da

    , (B33)

    where ay 5 (4/3)pa*. Because qi and an are known atany time, the mean effective density of the particles can

    be diagnosed from this relationship.

    e. Fall speeds and bulk ventilation coefficients

    A difficulty in estimating bulk fall speeds and venti-

    lation effects is that the Best number and the Reynolds

    number, both of which depend on size, are used to

    choose size-dependent coefficients. The main assump-

    tionmade here is that the distribution-averaged forms of

    each number can be used to choose the coefficients. The

    distribution-averaged Best number is

    X51

    Ni

    ðX(a)n(a) da5 xna

    bn

    nG(n1 bn)

    G(n), (B34)

    where xn and bn are defined in Eq. (A2) in appendix A.

    The mean value of X is used to determine the values of

    the coefficients (am and bm) needed to find the Reynolds

    number [Eq. (A3)] in the equation for the fall speed [Eq.

    (A1)].

    With the Reynolds number coefficients determined,

    the average fall speeds are computed by integrating Eqs.

    (A1) and (A3) over the a-axis distribution using the mo-

    ments defined in section a of appendix B. The number-

    weighted fall speed is defined as

    yt 51

    Ni

    ðNReharaL

    n(a) da5NRe(an)haraL(an)

    G(n1 bnbm 2bl)

    G(n),

    (B35)

    which is Eq. (17) in the main text. The mass-weighted

    fall speed is

    ytm5

    ðNReharaL

    m(a)n(a) daðm(a)n(a) da

    5NRe(an)haraL(an)

    G(n1bnbm2 bl 1 21 d*)

    G(n1 21 d*), (B36)

    where m(a) is defined in Eq. (7). Similarly, the length-

    weighted fall speed can be found from

    ytl 5

    ðNReharaL

    L(a)n(a) daðL(a)n(a) da

    5NRe(an)haraL(an)

    G(n1 bnbm)

    G(n1 bl),

    (B37)

    with L(a) defined in Eq. (A1).

    Similarly to the Best number, the Reynolds number

    is needed to choose values of the power-law co-

    efficients (gy and gt) that appear in the equations for

    ventilation [Eq. (A4)]. To pick these coefficients, the

    distribution-averaged Reynolds number [Eq. (A3)]

    is used:

    NRe 51

    Ni

    ðNRe(a)n(a) da5 amx

    bm

    n abnbm

    nG(n1 bnbm)

    G(n),

    (B38)

    where all variables are defined in appendix A.

    Once gy and gt are determined, bulk forms of the

    ventilation coefficients can be derived. However, the

    ventilation coefficients appear in the growth equation

    inside the functionGi in such a way that integration over

    the size spectrum is not possible. An alternative method

    is to begin with the equations for vapor and thermal

    diffusion [Lamb and Verlinde 2011, their Eq. (8.6),

    p. 324] prior to their combination into one final equation.

    Using vapor diffusion as an example, the distribution-

    averaged growth rate is

    dm

    dt5

    1

    Ni

    ðdm

    dtn(a) da5

    1

    Ni

    ð4pCDy fyDryn(a) da

    5 4pCDy fyDry , (B39)

    where Dry 5 ry,‘ 2 ry,i is the excess vapor density overthe crystal with ry,‘ and ry,i as the respective environ-

    mental vapor and equilibrium ice vapor densities, C is

    the distribution-averaged capacitance [Eq. (B14)], and

    fy is the bulk ventilation coefficient for vapor. A similar

    integral appears for thermal diffusion, and so the bulk

    ventilation coefficients are

    362 JOURNAL OF THE ATMOSPHER IC SC IENCES VOLUME 70

  • fy 5

    ðCfy(a)n(a) dað

    Cn(a) da

    and ft 5

    ðCft(a)n(a) daðCn(a) da

    . (B40)

    Once the capacitance-weighted ventilation coefficients

    are known, they can be multiplied by the vapor diffu-

    sivity and the thermal conductivity as an estimate of the

    overall effects of ventilation on the growth rate of

    a distribution of particles. The above integrals are solved

    using solutions to the moments of the a-axis distribution

    [e.g., Eq. (B1)] and by combining Eqs. (A4), (B7), and

    (B9) and rearranging to arrive at

    fy 5 by11 by2 N1/3Sc NRe(an)

    1/2h ig

    yf cap,y ,

    ft 5 bt11 bt2 N1/3Pr NRe(an)

    1/2h ig

    yf cap,t . (B41)

    These are Eq. (18) in the main text where f cap,y and f cap,tare factors that depend on the capacitance:

    f cap,y 51

    C

    �c1a

    d1

    nG(n1by 1 d1)

    G(n)1 c2a

    d2

    nG(n1 by 1 d2)

    G(n)

    �,

    f cap,t 51

    C

    �c1a

    d1

    nG(n1 bt 1d1)

    G(n)1 c2a

    d2

    nG(n1 bt 1d2)

    G(n)

    �,

    (B42)

    with by 5 (bnbmgy)/2 and bt 5 (bnbmgt)/2.It is possible to recast the ventilation coefficients for

    each axis so that they depend only on the a-axis length

    by using Eqs. (6), (8), and (A5):

    fya5 by11 by2Ngy/3

    Sh Ngy/2

    Re [a21/3* a

    1/3(12d*)]gy /2,

    fyc 5 by11 by2Ngy/3

    Sh Ngy/2

    Re [a2/3* a

    2/3(d*21)

    ]gy /2 , (B43)

    where all of the above coefficients are given in appendix

    A. For consistency, the axis-dependent ventilation co-

    efficients are averaged over the a-axis distribution in the

    same fashion as the total ventilation coefficient above;

    namely,

    f ya 5

    ðCfya(a)n(a) dað

    Cn(a) da

    f yc 5

    ðCfyc(a)n(a) dað

    Cn(a) da

    . (B44)

    Carrying out the integration and rearranging terms as

    was done for Eq. (18) produces

    f ya5by11by2 N1/3Sc NRe(an)

    1/2gya21/3* a

    1/3(12d*)n

    gy/2fcap,a,

    ihih

    f yc5by11by2 N1/3Sc NRe(an)

    1/2h ig

    ygya2/3* a

    2/3(d*21)n

    h igy/2fcap,c ,

    (B45)

    and this is Eq. (19) in the main text. The terms f cap,a and

    f cap,c are factors that depend on the capacitance:

    f cap,a51

    C

    �c1a

    d1

    nG(n1 bya1 d1)

    G(n)1 c2a

    d2

    nG(n1 bya1 d2)

    G(n)

    �,

    f cap,c 51

    C

    �c1a

    d1

    nG(n1 byc 1d1)

    G(n)1 c2a

    d2

    nG(n1 byc 1 d2)

    G(n)

    �.

    (B46)

    The coefficients bya and byc are defined as

    bya5gy2

    �bnbm 1

    2

    3(d*2 1)

    byc 5gy2

    �bnbm 1

    1

    3(12 d*)

    �. (B47)

    REFERENCES

    Avramov, A., and J. Y. Harrington, 2010: Influence of parame-

    terized ice habit on simulated mixed phase Arctic clouds.

    J. Geophys. Res., 115, D03205, doi:10.1029/2009JD012108.

    ——, andCoauthors, 2011: Towards ice formation closure inArctic

    mixed-phase boundary layer clouds during ISDAC. J. Geo-

    phys. Res., 116, D00T08, doi:10.1029/2011JD015910.

    Bailey, M., and J. Hallett, 2002: Nucleation effects on the habit of

    vapour grown ice crystals from 2188 to 2428C. Quart. J. Roy.Meteor. Soc., 128, 1461–1483.

    ——, and ——, 2004: Growth rate and habits of ice crystals be-

    tween 2208 and 2708C. J. Atmos. Sci., 61, 514–554.——, and ——, 2009: A comprehensive habit diagram for at-

    mospheric ice crystals: Confirmation from the laboratory,

    AIRS II, and other field studies. J. Atmos. Sci., 66, 2888–

    2899.

    Bohm, J., 1992: A general hydrodynamic theory for mixed-phase

    microphysics. Part I: Drag and fall speed of hydrometeors.

    Atmos. Res., 27, 253–274.

    Chen, J.-P., and D. Lamb, 1994: The theoretical basis for the pa-

    rameterization of ice crystal habits: Growth by vapor de-

    position. J. Atmos. Sci., 51, 1206–1221.

    ——, and ——, 1999: Simulation of cloud microphysical and

    chemical processes using a multicomponent framework. Part

    II: Microphysical evolution of a wintertime orographic cloud.

    J. Atmos. Sci., 56, 2293–2312.

    Fridlind, A., A. Ackerman, G. McFarquhar, G. Zhang, M. Poellot,

    P. DeMott, A. Prenni, and A. Heymsfield, 2007: Ice properties

    of single-layer stratocumulus during the Mixed-Phase Arctic

    FEBRUARY 2013 HARR INGTON ET AL . 363

  • Cloud Experiment: 2. Model results. J. Geophys. Res., 112,

    D24202, doi:10.1029/2007JD008646.

    Fukuta, N., and T. Takahashi, 1999: The growth of atmospheric ice

    crystals: A summary of findings in vertical supercooled cloud

    tunnel studies. J. Atmos. Sci., 56, 1963–1979.

    Harrington, J. Y., M. P. Meyers, R. L. Walko, and W. R. Cotton,

    1995: Parameterization of ice crystal conversion processes due

    to vapor deposition for mesoscale models using double-moment

    basis functions. Part I: Basic formulation and parcel model re-

    sults. J. Atmos. Sci., 52, 4344–4366.

    ——, K. Sulia, andH.Morrison, 2013: Amethod for adaptive habit

    prediction in bulk microphysical models. Part II: Parcel model

    corroboration. J. Atmos. Sci., 70, 365–376.

    Hashino, T., and G. J. Tripoli, 2007: The Spectral Ice Habit Pre-

    diction System (SHIPS). Part I: Model description and sim-

    ulation of the vapor deposition process. J. Atmos. Sci., 64,

    2210–2237.

    ——, and ——, 2008: The Spectral Ice Habit Prediction System

    (SHIPS). Part II: Simulation of nucleation and depositional

    growth of polycrystals. J. Atmos. Sci., 65, 3071–3094.

    Lamb, D., and W. Scott, 1972: Linear growth rates of ice crystals

    grown from the vapor phase. J. Cryst. Growth, 12, 21–31.

    ——, and J. Verlinde, 2011: The Physics and Chemistry of Clouds.

    Cambridge University Press, 584 pp.

    Libbrecht, K., 2003: Growth rates of the principal facets of ice

    between 2108C and 2408C. J. Cryst. Growth, 247, 530–540.——, 2005: The physics of snow crystals. Rep. Prog. Phys., 65, 855–

    895.

    Lin, Y., R. Farley, and H. Orville, 1983: Bulk parameterization of

    the snow field in a cloud model. J. Climate Appl. Meteor., 22,1065–1092.

    López, M. L., and E. E. Ávila, 2012: Deformation of frozen drop-

    lets formed at 2408C. Geophys. Res. Lett., 39, L01805,doi:10.1029/2011GL050185.

    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, 3–39.Milbrandt, J. A., and M. K. Yau, 2005: A multimoment bulk

    microphysics parameterization. Part II: A proposed three-

    moment closure and scheme description. J. Atmos. Sci., 62,

    3065–3081.

    Mitchell, D. L., 1996: Use of mass- and area-dimensional power

    laws for determining precipitation particle terminal velocities.

    J. Atmos. Sci., 53, 1710–1723.

    ——, R. Zhang, and R. L. Pitter, 1990: Mass-dimensional relation-

    ships for ice particles and the influence of riming on snowfall

    rates. J. Appl. Meteor., 29, 153–163.

    Morrison, H., and W. Grabowski, 2008: A novel approach for

    representing ice microphysics in models: Description and tests

    using a kinematic framework. J. Atmos. Sci., 65, 1528–1548.

    ——, and ——, 2010: An improved representation of rimed snow

    and conversion to graupel in a mutlicomponent bin micro-

    physics scheme. J. Atmos. Sci., 67, 1337–1360.

    Nelson, J., 1994: A theoretical study of ice crystal growth in the

    atmosphere. Ph.D. dissertation, University of Washington,

    183 pp.

    ——, and C. Knight, 1998: Snow crystal habit changes explained by

    layer nucleation. J. Atmos. Sci., 55, 1452–1465.

    Pruppacher, H. R., and J. D. Klett, 1997: Microphysics of Clouds

    and Precipitation. Kluwer Academic Publishers, 954 pp.

    Reisin, T., Z. Levin, and S. Tzivion, 1996: Rain production in

    convective clouds as simulated in an axisymmetric model

    with detailed microphysics. Part I: Description of the model.

    J. Atmos. Sci., 53, 497–519.

    Reisner, J., R. Rasmussen, and R. Bruintjes, 1998: Explicit fore-

    casting of supercooled liquid in winter storms using the MM5

    mesoscale model.Quart. J. Roy. Meteor. Soc., 124, 1071–1107.

    Sheridan, L. M., J. Y. Harrington, D. Lamb, and K. Sulia, 2009:

    Influence of ice crystal aspect ratio on the evolution of ice size

    spectra during vapor depositional growth. J. Atmos. Sci., 66,3732–3743.

    Sulia, K. J., and J. Y. Harrington, 2011: Ice aspect ratio influences

    on mixed-phase clouds: Impacts on phase partitioning in

    parcel models. J. Geophys. Res., 116, D21309, doi:10.1029/

    2011JD016298.

    Thompson, G., R. Rasmussen, and K. Manning, 2004: Explicit

    forecasts of winter precipitation using an improved bulk mi-

    crophysics scheme. Part I: Description and sensitivity analysis.

    Mon. Wea. Rev., 132, 519–542.

    ——, P. Field, R. Rasmussen, and W. Hall, 2008: Explicit forecasts

    of winter precipitation using an improved bulk microphysics

    scheme. Part II: Implementation of a new snow parameteri-

    zation. Mon. Wea. Rev., 136, 5095–5115.

    Walko, R. L., W. R. Cotton, M. P. Meyers, and J. Y. Harrington,

    1995: New RAMS cloud microphysics parameterization. Part

    I: The single moment scheme. Atmos. Res., 38, 29–62.

    ——, ——, G. Feingold, and B. Stevens, 2000: Efficient computa-

    tion of vapor and heat diffusion between hydrometeors in

    a numerical model. Atmos. Res., 53, 171–183.

    Woods, C., M. Stoelinga, and J. Locatelli, 2007: The IMPROVE-1

    storm of 1–2 February 2001. Part III: Sensitivity of a mesoscale

    model simulation to the representation of snow particle types

    and testing of a bulk microphysical scheme with snow habit

    prediction. J. Atmos. Sci., 64, 3927–3948.

    364 JOURNAL OF THE ATMOSPHER IC SC IENCES VOLUME 70