1: \documentclass[12pt,preprint]{aastex}
2: \newcommand{\vdag}{(v)^\dagger}
3: \newcommand{\myemail}{Paul.R.Estrada@nasa.gov}
4:
5: \shorttitle{Moments Method}
6: \shortauthors{Estrada and Cuzzi}
7:
8: \begin{document}
9:
10: \title{Solving the Coagulation Equation by the Moments Method}
11:
12: \author{P. R. Estrada}
13: \affil{SETI Institute, 515 N. Whisman Rd., Mountain View, CA 94043}
14:
15: \and
16:
17: \author{J. N. Cuzzi}
18: \affil{NASA Ames Research Center, MS 245-3, Moffett Field, CA 94035}
19:
20:
21: \begin{abstract}
22: We demonstrate an approach to solving the coagulation equation that involves
23: using a finite number of moments of the particle size distribution. This
24: approach is particularly useful when only
25: general properties of the distribution, and their time evolution, are needed.
26: The numerical solution to the integro-differential Smoluchowski coagulation
27: equation at every time step, for every particle size, and at every spatial
28: location is
29: computationally expensive, and serves as the primary bottleneck in running
30: evolutionary models over long periods of time. The advantage of using
31: the
32: moments method comes in the computational time savings gained from only
33: tracking the time rate of change of the moments, as opposed to tracking the
34: entire mass histogram which can contain hundreds or thousands of
35: bins depending on the desired accuracy. The collision kernels of the
36: coagulation equation contain all the necessary information about particle
37: relative velocities, cross-sections, and sticking coefficients. We show how
38: arbitrary collision kernels may be treated. We discuss particle relative
39: velocities in both turbulent and non-turbulent regimes.
40: We present examples of this approach
41: that utilize different collision kernels and find good agreement between
42: the moment solutions and the moments as calculated from direct integration
43: of the coagulation equation. As practical applications, we demonstrate
44: how the moments method can be used to track the evolving opacity, and
45: also indicate how one may incorporate porous particles.
46: \end{abstract}
47:
48:
49: \keywords{accretion, accretion disks --- methods: numerical --- radiative
50: transfer --- solar system: formation}
51:
52:
53: \section{Introduction}
54:
55: The conditions under which planetary formation is initiated in the
56: protoplanetary disk remain poorly understood despite several decades of
57: astrophysical research. Furthermore, the continued
58: discovery of extrasolar planets (over two hundred currently) further
59: emphasizes the need to understand the formation of not only our own solar
60: system, but of these numerous strange and diverse systems as well.
61: At the forefront still remains the basic question of
62: how growth occurs from dust to planet. The key properties of protoplanetary
63: nebulae remain controversial, yet grain growth from dust to larger
64: agglomerates
65: ($\sim $ cm-m), and from planetesimals ($\gtrsim$ km) to planets must have
66: occurred in some manner. While numerous studies of $N$-body dynamics have
67: addressed the growth of terrestrial planets from asteroid-size planetesimals
68: which are largely decoupled from the gas \citep[e.g.,][]{lei05,bro06},
69: it is the transition from agglomerates to
70: planetesimals which continues to provide the major stumbling block in
71: planetary origins \citep{wei97,wei02,wei04,wei93,ste97,dul05,cuz06,dom07}.
72:
73: Grain growth begins with sticking of sub-micron sized grains, which are
74: dynamically coupled to the nebula gas and collide at
75: low relative velocities which are
76: size-dependent \citep{wei93}. These relative velocities can be caused
77: by a variety of mechanisms, such as Brownian motion for the smallest grains,
78: differences in coupling to eddies in a
79: turbulent nebula \citep{vol80,wei84,mar91,wei93,cuz03,orm07},
80: and/or vertical, radial, and azimuthal drift driven by global gas pressure
81: gradients
82: \citep{nak86,wei97}. For the
83: smallest sized grains and aggregates, sticking is caused by weak van der
84: Waals interaction forces \citep[although stronger forces may also act; see,
85: e.g.,][]{dom07}. This sticking forms larger and larger aggregates
86: until particles grow large enough to decouple from the gas and
87: settle to the midplane where they may, barring strong turbulence
88: \citep[although see][for an exception]{joh07}, grow into
89: larger objects, even planetesimals \citep{saf69,wei93,cuz93,wei97}.
90:
91: There is considerable observational evidence supporting dust growth
92: in protoplanetary disks, at least from sub-micron to millimeter scales, based
93: on observations of the dust continuum emission originating from protoplanetary
94: disks. The wavelength dependence of dust emission from IR to radio wavelengths
95: suggests, in many cases, that typical grain sizes are in the millimeter range,
96: which is orders of magnitude larger than interstellar grains
97: \citep{bec91,dul05,dom07}.
98: Measurements of the contrast of the Orion trapezium
99: silicate emission feature at 10 $\mu$m relative to the local continuum implies
100: an increase in size from interstellar dust (sub-micron) to micron sizes in the
101: disk photospheres \citep{van03,pry03,kes06}.
102:
103: Theoretical approaches to the study of dust coagulation normally involve
104: solving the cumbersome collisional coagulation equation (see Eq. 1) at each
105: spatial location ($R,\phi,Z$) in the disk.
106: The collisional kernel $K(m,m^\prime)$ for particles of mass $m$ and $m^\prime$
107: usually involves some sort of sticking efficiency, and particle cross-sectional
108: area, in addition to the particle-particle
109: relative velocities. Although the {\it systematic} particle-particle
110: velocities that arise from
111: gas pressure gradients \citep{nak86} are
112: somewhat easy to incorporate into the collisional kernel,
113: relative velocities in turbulence are more complicated.
114: If one assumes that nebula turbulence follows a Kolmogorov inertial range
115: \citep{vol80} in which energy
116: cascades from the largest, slowly rotating eddies, down to some minimum
117: lengthscale where the intrinsic molecular viscosity can dissipate the
118: macroscopic gas motions (and turbulence ceases), one can obtain closed-form
119: expressions for the particle-particle, and particle-gas
120: turbulent relative velocities \citep{cuz03}, even for particles
121: of different sizes \citep{orm07}. Numerical
122: models of particle growth in the inner regions of protoplanetary disks
123: indicate that inclusion of these turbulence-induced relative
124: velocities is essential \citep{dul05}, as it provides a
125: source of energetic collisions that, in the presence of fragmentation
126: and destruction, leads to the
127: ongoing existence of small grains even after several million years.
128:
129: A full scale solution to the problem of dust coagulation for a size spectrum
130: at every vertical and radial location in the protoplanetary disk, as a
131: function of time,
132: is thus prohibitive even with today's computational capabilities
133: \citep{wei02,wei04}. So it is
134: advantageous to find alternative methods of extracting information from the
135: coagulation process, such as time scales of growth, the largest particle
136: size, the size that dominates the area (radiative transfer), and the size
137: that dominates
138: the mass (migration and redistribution) without having to track the behavior
139: of the entire mass distribution.
140:
141: In
142: this paper, we describe a method that utilizes the moments of the coagulation
143: equation to accomplish this task, in particular when
144: the functional form of the mass distribution can be assumed. In {\S} 2,
145: we derive the
146: moment equations from the coagulation equation, and compare
147: direct integration of the coagulation equation with solutions using
148: the moments approach in the case of a simple turbulent coagulation kernel.
149: We also present
150: two different approaches to obtaining solutions using a realistic kernel,
151: when the form of the size distribution is assumed to be a powerlaw.
152: In {\S} 3, we evaluate the different approaches developed in {\S} 2
153: for realistic kernels by numerical integration of the moments equations
154: and compare these moments to those obtained by direct integration of the
155: coagulation equation.
156: We also
157: compare the moments solution to an alternate analytical approach by
158: \citet{gar07}. In {\S} 4, as a demonstration of the practical application
159: of
160: the moments method, we calculate the wavelength-independent opacity, and
161: discuss generalization to wavelength-dependent Planck or Rosseland
162: mean opacities. In {\S} 5, we indicate how porosity may be included.
163: In {\S} 6, we present our conclusions.
164:
165: \section{Solution Techniques Using the Moments Method}
166:
167: The moments method approach to modeling particle growth allows an attractive
168: alternative to solving the full coagulation equation when it is only necessary
169: to know a particular property, or properties of the evolving particle size
170: distribution. The coagulation equation \citep{smo16} in its
171: integro-differential form is given by
172:
173: \begin{eqnarray}
174: \frac{df(m,t)}{dt} = \frac{1}{2}\int_{0}^{m} K(m-m^{\prime},
175: m^{\prime}) f(m-m^{\prime},t) f(m^{\prime},t) dm^{\prime} - \nonumber \\
176: \int_{0}^{\infty}
177: K(m,m^{\prime})f(m,t) f(m^{\prime},t) dm^{\prime} \,\,\,+ \,\,\,
178: \frac{df}{dt}\Big|_{+}\,\,\,+\,\,\, \frac{df}{dt}\Big|_{-}\,\,,
179: \end{eqnarray}
180:
181: \noindent
182: where $f(m,t)$ is the particle number density per unit mass at mass $m$, and
183: $K(m,m^{\prime})$ is the collision kernel connecting the properties of
184: interacting
185: masses $m$ and $m^\prime$, which typically takes into account mutual particle
186: cross section $\sigma(m,m^\prime)$, relative velocity $\Delta V(m,m^\prime)$,
187: and a particle sticking efficiency $S(m,m^\prime)$. The last two terms on
188: the RHS of Eq. (1) represent sources and sinks such as particle
189: erosion, fragmentation and subsequent redistribution of the fragmented
190: population, and gravitational growth. Although we will not specifically
191: address these mechanisms, we will indicate how they may be treated, and
192: save implementation for a forthcoming paper.
193:
194: The
195: motivation behind using the moments method is to avoid the computational cost
196: inherent in Eq. (1) which entails solving the convolution integral (the first
197: term on the RHS of Eq. 1) at every location, and at every time step.
198: Depending on the functional form of the kernel, the second integral on the
199: RHS may also need to be integrated repeatedly. Given
200: that the typical mass spectrum may involve $10^2-10^3$ bins to acquire the
201: desired accuracy, the computational burden required becomes a detriment to any
202: study involving a wide range of parameter space. The
203: so-called brute force solution of the coagulation equation thus becomes the
204: primary bottleneck in running 2D evolutionary models over extended periods of
205: time \citep{wei04,dul05}.
206:
207: We define the $p$-th moment $M_p$ of the distribution, where $p$ need not be
208: an integer, as
209:
210: \begin{equation}
211: M_p = \int_{0}^{\infty} m^p f(m,t) dm,
212: \end{equation}
213:
214: \noindent
215: where the units of $f(m,t)$ are such that $M_0=\int f(m,t)\,dm$ represents
216: the total number
217: density of particles, and $M_1 = \int mf(m,t)\, dm = \rho$ is the
218: total volume mass density
219: of solids. The essence of the moments method is as follows. We multiply both
220: sides of the coagulation equation (Eq. 1) by $m^k$, where $k$ is an integer,
221: and then integrate both sides over mass $m$:
222:
223: \begin{eqnarray}
224: \frac{dM_k}{dt} = \int_{0}^{\infty} m^k \frac{df}{dt} dm =
225: \frac{1}{2} \int_{0}^{\infty} m^k\,dm \int_{0}^{m} K(m-m^{\prime},
226: m^{\prime}) f(m-m^{\prime},t) f(m^{\prime},t) dm^{\prime} - \nonumber \\
227: \int_{0}^{\infty} m^k f(m,t)\,dm \int_{0}^{\infty}
228: K(m,m^{\prime}) f(m^{\prime},t) dm^{\prime}.
229: \end{eqnarray}
230:
231: \noindent
232: We introduce a step function $H(m-m^\prime)$, such that $H = 0$ for
233: $m -m^\prime < 0$ and $H = 1$ otherwise, to extend the limits of the
234: integral over $m^\prime$ from ($0,m$) to ($0,\infty$). The convolution
235: integral (the first integral on the RHS of Eq. 3) then becomes
236:
237: \begin{eqnarray}
238: \frac{1}{2}\int_{0}^{\infty} m^k\,dm \int_{0}^{\infty}
239: H(m-m^\prime) K(m-m^\prime,m^\prime) f(m-m^\prime,t) f(m^\prime,t)\,
240: dm^\prime = \nonumber \\
241: \frac{1}{2} \int_{0}^{\infty} f(m^\prime,t)\,dm^\prime \int_{0}^{\infty}
242: (u+m^\prime)^k K(u,m^\prime) f(u,t)\,du,
243: \end{eqnarray}
244:
245: \noindent
246: where on the RHS of Eq. (4) we have switched the order of integration and
247: made the substitution $u = m - m^\prime$. The RHS side of Eq. (4) may then be
248: combined with the last double integral in Eq. (3) to yield the set of
249: ordinary differential equations (ODEs)
250: for the integer moments \citep{mar01}:
251:
252: \begin{equation}
253: \frac{dM_k}{dt} = \int_{0}^{\infty} \int_{0}^{\infty} \left[\frac{1}{2}
254: (m + m^\prime)^k - m^k\right] K(m,m^\prime) f(m,t) f(m^\prime,t) dm dm^\prime,
255: \end{equation}
256:
257: \noindent
258: where we have substituted $m$ for $u$ with no loss of generality.
259: Depending on the mass dependence of the kernel (as illustrated below)
260: the right hand side can readily be
261: expressed as products of moments of order $k$ or less, leading to a closed
262: system of equations which may be solved using standard techniques.
263: In this definition of the coagulation equation in which it is assumed there
264: are no sources and sinks, the first
265: moment $M_1 \equiv \rho$ is constant in time, that is
266: $dM_1/dt = 0$\footnote{We note that under realistic protoplanetary nebula
267: conditions, $\rho$ will not be constant due to, e.g., size-dependent advection
268: terms in the equations. Such effects can be treated separately from the
269: ``coagulation'' step. See {\S} 2.2.2.}.
270:
271: Exact solutions to the coagulation equation have been obtained for
272: some specific choices
273: of the kernel \citep[e.g.,][]{smo16,tru71}, the most simple
274: being that of constant kernel $K(m,m^\prime) = \beta_0$. Although the exact
275: solution for $f(m,t)$ cannot be obtained from the moments equations, the
276: time rate of
277: change of the zeroth moment ($k=0$), can easily be obtained from Eq. (5)
278: which reduces to $dM_0/dt = - (1/2)\beta_0M_0^2$ and has the trivial solution
279: \citep[see, e.g.,][]{sil79}
280:
281: \begin{equation}
282: M_0(t) = \frac{M_0(0)}{1 + (1/2)\beta_0M_0(0)t},
283: \end{equation}
284:
285: \noindent
286: independent of the initial choice of distribution $f(m,0)$. Likewise, the
287: ODE for $M_2$, which can be associated with the density-weighted mean particle
288: size ($\left<m\right> = M_2/M_1$), yields the simple solution
289: $M_2(t)=M_2(0) + \beta_0\rho^2t$. Despite not
290: knowing the exact solution, we are able
291: to understand the behavior of general properties of the
292: mass distribution with time through the moments equation without
293: tracking the behavior of the full mass spectrum. Thus,
294: if it is only desired to know, for example, the time evolution of the
295: particle representing most of the {\it area} (first moment) or most of
296: the {\it mass} (second moment) of the distribution, then the advantage of
297: the moments method becomes clear. A small
298: number of moments is all that is necessary to determine the behavior of the
299: system. In particular, we will be interested in the size of the largest
300: particle $m_L(t)$ in the entire mass distribution as a function of time
301: (which we show below can be computed from the integer moments),
302: because these are usually the most rapidly drifting and most violently
303: colliding particles \citep{cuz06}.
304:
305:
306: \subsection{Example: Saffman and Turner Turbulent Coagulation Kernel}
307:
308: We can illustrate the moments method approach using a very simple turbulent
309: coagulation kernel
310: where $K(m,m^\prime) = \gamma_0 (m^{1/3} + m^{\prime 1/3})^3$ \citep{saf56},
311: and no sources or sinks. Physically, this represents the
312: product of $(r + r^{\prime})^2$ for
313: area, and $(r + r^{\prime})$ for relative velocity of two particles in a
314: laminar shear flowfield. If a sticking coefficient were desired then we
315: would specifically have $K(m,m^\prime) = \gamma_0(m^{1/3}+m^{\prime 1/3})^3
316: S(m,m^\prime)$ where $0\leq S(m,m^\prime) \leq 1$. Here, $\gamma_0$ is a
317: constant that depends on the Reynolds number of the gas flow.
318: Inserting this kernel into Eq. (5), we find the set of equations
319:
320: \begin{eqnarray}
321: \frac{dM_0}{dt} = -\gamma_0(M_0M_1 + 3M_{1/3}M_{2/3}), \,\,\,\,\,
322: \frac{dM_1}{dt} = 0 \nonumber \\
323: \frac{dM_2}{dt} = 2\gamma_0(M_2M_1 + 3M_{4/3}M_{5/3}).
324: \end{eqnarray}
325:
326:
327: \noindent
328: We note that the physical derivation of kernels in terms of particle radius $r$
329: leads to {\it fractional}
330: moments in terms of $m$. These fractional moments look complicated, but can
331: be solved for by simple
332: interpolation using Lagrange polynomials in terms of the more familiar integer
333: moments \citep{log79,pre92,mar01}.
334: Thus, any
335: fractional moment $M_p$ may be expressed compactly in the normalized form
336: $M^\prime_p(t) = M_p(t)/M_p(0)$:
337:
338: \begin{equation}
339: M^\prime_p(t) = \prod_{j=k}^{k+n} \left[M^\prime_j(t)\right]^{L^n_j(p)},
340: \end{equation}
341:
342: \noindent
343: where $n$ is the number of integer moments, $k = 0,1,...,n$,
344: and the exponent $L^n_j$ is defined as
345:
346: \begin{equation}
347: L^n_j(p) = \frac{1}{n!}{\prod}_{n+1}(p)\frac{(-1)^{n-j}C^j_n}{p-j},
348: \end{equation}
349:
350: \noindent
351: with ${\prod}_{n+1}(p) = p(p-1)...(p-n)$, and the $C^j_n$
352: are the binomial coefficients $n!/j!(n-j)!$. The
353: important thing to
354: note here is that the order of the moments must remain less than or equal
355: to the largest moment in order for the system to remain closed. In general,
356: this will be true for realistic collision kernels (see {\S} 2.2).
357:
358: To test the accuracy of the moments method, we integrated equations (7) using a
359: fourth order Runge-Kutta scheme, and compared the results to a brute force
360: integration of the coagulation equation (Eq. 1). For the latter, we
361: integrated the distribution function $f(m,t)$ at each timestep
362: to determine the moments of the distribution $M_0$, $M_1$, and $M_2$. We chose
363: $f(m,0) = c_0m^{-q}$ as our initial distribution for simplicity. Since the
364: form of the mass distribution is only assumed at $t=0$, we consider this to
365: be an example of an implicit assumption (see {\S} 2.2.2).
366: The results of this calculation
367: are presented in Figure 1. We have used two different resolutions for the
368: brute force calculation in order to demonstrate how the higher resolution
369: (and much more computationally expensive) case (solid symbols) approaches the
370: moments approach solution. Notice that the first moment $M_1 = \rho$ remains
371: constant as expected. Given that
372: general numerical errors can arise from the finite mass grid and coarse
373: timesteps ($\Delta t = 10$ years) used for
374: the calculation, and that systematic errors may be
375: introduced by the Lagrangian interpolation, the fit is quite good.
376: It is important
377: to point out here that while
378: the coagulation calculation required as many as $20-30$ hours of CPU time to
379: complete on a 2 GHz machine, the moments calculation of Eq. (7) is
380: essentially instantaneous.
381:
382:
383: \subsection{Realistic Coagulation Kernels}
384:
385: The Saffman-Turner turbulent coagulation kernel we used as an example in
386: {\S} 2.1
387: is simple to utilize and is expressible explicitly in powers of the mass $m$.
388: In practice, however, the realistic coagulation kernels that we will be using
389: will be more complicated than this simple example. A realistic
390: coagulation kernel will be, at the very least, a product of a mutual cross
391: section $\pi (r + r^\prime)^2$ and a relative velocity
392:
393: \begin{equation}
394: \Delta V(m,m^\prime) = \sqrt{ (U - U^\prime)^2 + (V - V^\prime)^2 +
395: (W - W^\prime)^2 + v_T^2(m,m^\prime)},
396: \end{equation}
397:
398: \noindent
399: where $U(m)$, $V(m)$, and $W(m)$ are the $x,y,z$ systematic (pressure
400: gradient driven) particle
401: velocities, and $v_T$ is the turbulent velocity contribution
402: which in general is not separable into distinct functions of
403: $m$ and $m^\prime$ \citep{cuz03,orm07}. The
404: systematic velocities,
405: which are derived for a discretized particle size distribution in the
406: appendix, are complicated functions of the particle size through the
407: stopping time $t_s$,
408:
409: \begin{equation}
410: t_s = \frac{m\Delta V_{pg}}{F_D},
411: \end{equation}
412:
413: \noindent
414: where $\Delta V_{pg} = |\vec{\bf{V}} - \vec{\bf{v}}|$ is the relative
415: velocity between the particle and the gas, and
416: $F_D$ is the drag force on the particle of size $r$ which depends on the size
417: of the particle relative to the mean free path $\lambda$ of the gas molecule
418: \citep[e.g., see][]{cuz93}. Depending on whether $r \gtrsim
419: \lambda$ or $r \lesssim \lambda$ determines whether the stopping time itself
420: depends on the instantaneous relative velocity $\Delta V_{pg}$
421: between the particle and the gas (Stokes regime) or does not (Epstein regime).
422: In the latter case, calculation of the systematic velocities ($U,V,W$) and
423: gas velocities ($u,v,w$) for a particle size distribution is
424: straightforward. In the former case, iterations are required to correctly
425: calculate $\Delta V_{pg}$ (see appendix).
426:
427: The turbulent velocities are less straightforward to implement than their
428: systematic counterparts because of the different coupling that exists
429: between particles and eddies of different sizes (and thus stopping and decay
430: times, respectively). It has not been until very recently that closed form
431: expressions for the turbulence-induced velocities (for a particle size
432: distribution) were derived \citep{orm07}. These expressions can be written
433: separately for the so-called ``class I'' and ``class II''
434: eddies\footnote{The concept of eddy classes were introduced by \citet{vol80}
435: to distinguish between slowly and rapidly varying eddies. Class I eddies
436: are defined as those in which the eddy fluctuates slowly enough that a
437: particle's
438: stopping time $t_s$ is much shorter than the eddy decay time and the time
439: it takes
440: to cross the eddy; thus, they will align themselves with the gas motions of
441: the eddy prior to its decay.
442: Class II eddies then are defined as ones in which the eddy decay time is fast
443: compared to the particles $t_s$, and thus only provide a
444: small perturbation to the particle motion.} as:
445:
446: \begin{equation}
447: \Delta V_I^2 = v_g^2\frac{{\rm{St}}_i-{\rm{St}}_j}{{\rm{St}}_i+{\rm{St}}_j}
448: \left(\frac{{\rm{St}}^2_i}{{\rm{St}}^*_{ij}
449: +{\rm{St}}_i} - \frac{{\rm{St}}^2_i}{1+{\rm{St}}_i} - \frac{{\rm{St}}^2_j}
450: {{\rm{St}}^*_{ij}+{\rm{St}}_j} +
451: \frac{{\rm{St}}^2_j}{1+{\rm{St}}_j}\right),
452: \end{equation}
453:
454: \begin{equation}
455: \Delta V^2_{II} = v_g^2\left( 2({\rm{St}}^*_{ij} - {\rm{Re}}^{-1/2}) +
456: \frac{{\rm{St}}^2_i}{{\rm{St}}_i+{\rm{St}}^*_{ij}} - \frac{{\rm{St}}^2_i}
457: {{\rm{St}}_i+
458: {\rm{Re}}^{-1/2}} + \frac{{\rm{St}}^2_j}{{\rm{St}}_j+{\rm{St}}^*_{ij}} -
459: \frac{{\rm{St}}^2_j}{{\rm{St}}_j+{\rm{Re}}^{-1/2}}\right),
460: \end{equation}
461:
462: \noindent
463: where $v_T^2 = \Delta V^2_I+\Delta V^2_{II}$, ${\rm{Re}}$ is the Reynolds
464: number, $v_g = \alpha^{1/2}c_g$ is the turbulent gas velocity with $c_g$ the
465: sound speed \citep[e.g.,][]{nak86,cuz06}, and the particle Stokes numbers
466: ${\rm{St}}_i = t_{{\rm{s}}i}/t_L$, where $t_L$ is the turnover time of the
467: largest eddy, typically
468: taken to be $\Omega^{-1}$. The boundary between class I and class II is
469: defined by the ``combined'' Stokes number
470: ${\rm{St}}^*_{ij} = {\rm{max}}({\rm{St}}^*_i,{\rm{St}}^*_j)$
471: and the values of the ${\rm{St}}^*_k$ are obtained from the
472: equation \citep{orm07}:
473:
474: \begin{equation}
475: \frac{2}{3}y_k^*(y_k^*-1)^2 - \frac{1}{1+y_k^*} = -\frac{{\rm{St}}_k}
476: {1+{\rm{St}}_k} + \frac{1}{{\rm{St}}_k}\frac{\Delta V^2_{pg}}{v^2_g},
477: \end{equation}
478:
479: \noindent
480: where $y^*_k = {\rm{St}}^*_k/{\rm{St}}_k$.
481: In our calculations, we will make use of these expressions when comparing
482: cases with turbulence-induced velocities.
483:
484: Given that the moments method requires that we
485: express the kernel in terms of fractional or integer moments, the problem
486: becomes expressing the relative velocity (Eq. 10) in a form that satisfies
487: this requirement. As an
488: example, the systematic velocities $U$, $V$, and $W$ can each be fit easily
489: by a finite series in fractional powers $p_i$ of $m$
490: as $U(m) = \sum^N_i a_i m^{p_i}$ where the
491: coefficients $a_i$ can be found by fitting $N$ points to the function
492: $U(m_l) = U_l$ ($l$ an integer). This leads to the system of equations
493:
494: \begin{equation}
495: U_l = \sum_{i=1}^N a_i m_l^{p_i}.
496: \end{equation}
497:
498: \noindent
499: The coefficients $a_i$ then
500: follow from matrix inversion. Similarly, one can find expressions
501: $V(m) = \sum_i^N b_i m^{p_i}$, and $W(m) = \sum_i^N c_i m^{p_i}$
502: so that we may
503: construct the laminar expression for $(\Delta V)^2$ in terms of the variables
504: $m$ and
505: $m^\prime$ by multiplying out the individual terms, e.g.,
506: $U(m)U(m^\prime) = \sum_{i,j} a_ia_j m^{p_i}m^{\prime p_j}$ and so on. We
507: find then that we can express the square of Eq. (10) as
508:
509: \begin{equation}
510: (\Delta V(m,m^\prime))^2 = \sum_{i,j} A_{ij} \left[m^{p_i+p_j} -
511: 2m^{p_i}m^{\prime p_j} + m^{\prime p_i+p_j}\right],
512: \end{equation}
513:
514: \noindent
515: where $A_{ij} = a_ia_j + b_ib_j + c_ic_j$, and $p_i+p_j\le 2$ for closure.
516: This determines the matrix of
517: coefficients $A_{ij}$ which operate on the finite power series in
518: $m$ and $m^\prime$. Although we cannot use the same approach for solving
519: for $\Delta V$ in
520: Eq. (10) because of the radical, we were motivated to express $\Delta V$ in
521: a similar form, i.e., as
522: $\Delta V(m,m^\prime) = \sum_{i,j} A^\prime_{ij} [m^{p_i+p_j} -
523: 2m^{p_i}m^{\prime p_j} + m^{\prime p_i+p_j}]$, and $p_i+p_j\le 1$.
524: The problem reduces to solving for the
525: coefficients $A^\prime_{ij}$ directly, by similar matrix
526: inversion techniques. Unfortunately, the number of points needed to obtain
527: an accurate representation of the two-variable function
528: $\Delta V(m,m^\prime)$ this way
529: exceeds the limitations of the inversion. In
530: practice we found that we could solve a system for $\Delta V$
531: with at most
532: $5-6$ points ($25-36$ matrix elements),
533: whereas Eq. (16) carries much less restriction in the
534: number of points needed because we could construct $(\Delta V)^2$
535: by the product of accurate representations of single variable functions.
536: Furthermore, due to the coupling evident in
537: the turbulence induced velocities (Equations 12 and 13), the turbulent
538: velocities are not separable functions of the masses $m$ and $m^\prime$,
539: further complicating the analysis.
540:
541: {\it Powerlaw assumption}: Fortunately, it turns out that, although
542: the direct calculation approach to $\Delta V$ described above would
543: be mathematically appealing in the sense that the moment equations would
544: remain implicit (i.e., the form of the mass distribution would only
545: be assumed at $t=0$), it is not necessary. In the next sections
546: we describe two alternative approaches to dealing with realistic coagulation
547: kernels that employ the moments method under the assumption that the
548: form of the mass distribution $f(m,t)$ is a powerlaw.
549: Such an assumption is
550: not entirely unfounded. A number of detailed models by
551: \citet{wei97,wei00,wei04}
552: have shown that powerlaw size distributions result, which have
553: nearly constant mass per decade radius to an upper limit $m_L(t)$ which
554: grows with time until the frustration limit of around a meter is reached, and
555: (under turbulent conditions, at least)
556: growth stalls \citep{cuz06}. Similar trends are found by \citet{dul05} in
557: which different assumptions about collisional ejecta are made. These authors
558: do find minor fluctuations in the distribution, but if one is primarily
559: interested in general properties of the distribution, such as
560: how the largest particle size changes with time, and not the fine
561: structure of the
562: mass distribution itself, the approach is quite advantageous. There are reasons
563: to believe that a powerlaw is a natural end-state, especially those
564: with equal mass per decade, because they have self-preserving properties for
565: the collision kernels of interest \citep{cuz06}.
566:
567: The significance of the upper mass cutoff $m_L$ depends on the assumed slope
568: of the powerlaw. In all real distributions, there will be a rapidly decreasing
569: abundance of particles for masses exceeding some threshold, even though the
570: abundance may not drop immediately to zero as in our assumed model. The rare,
571: extremely large particles might be of interest for some applications, but
572: our focus will be the particle size carrying most of the area, or most of
573: the mass (the first or second moments). For powerlaw distributions
574: $f(m) \propto m^{-q}$ with $q < 2$, $m_L$ is itself the mass of the particle
575: carrying most of the mass and is independent of the selection of a lower
576: particle size cutoff. For $q=2$, the distribution contains equal mass per
577: decade, and for $q>2$, most of the mass is in the small particles and the
578: value of $m_L$ depends on the lower particle size chosen. Most realistic
579: distributions, and those most commonly treated in the literature, have $q<2$;
580: here we assume $q=11/6$, a widely used fragmentation powerlaw. In this case,
581: a strict cutoff at $m_L$ represents a well-defined distribution with
582: easily-understood moments where most of the mass is at $m_L$ and the area
583: is nearly equally distributed per decade with a mass dependence $m^{-1/6}$.
584:
585: Although we do not include either imperfect sticking or fragmentation at this
586: stage in the model, we believe that the moment equations as expressed in
587: Eq. (5) will remain valid up to a "fragmentation barrier", which may be
588: defined as that size for which the typical disruption energy of a particle
589: is on the same order as the energy of identical colliding particles.
590: The fragmentation barrier will also depend on one's choice of
591: nebula parameters. This treatment (up to the fragmentation size) is
592: consistent with recent work by \citet{bra08} (their Fig. 13) which shows
593: a constant powerlaw mass distribution up to a cutoff size which then falls
594: off abruptly. This ``knee'' in the distribution represents that efficient
595: fragmentation size.
596:
597: Once $m_L$ reaches the fragmentation barrier, growth beyond this
598: stage would need to be treated in a different manner, for
599: example, by simple sweepup of small, less disruptive particles by large
600: ones \citep{cuz93}. Creation and disruption of these
601: particles can be handled as part of the source and sink terms in which
602: for example, disrupted particles are assumed to be fragmented back into
603: a powerlaw distribution (which is suggested by experimental evidence and
604: widely assumed by other modelers) as opposed to monomers. The model as
605: presented within this paper, however, should be useful for the early
606: stages of protoplanetary nebula particle growth relevant to spectral
607: energy distributions and MRI suppression. We leave the incorporation of
608: growth stages beyond the efficient fragmentation stage for a later paper.
609:
610: \subsubsection{Approach 1: Explicit Assumption}
611:
612: Motivated by our discussion of the previous section, if we assume that the
613: form of the particle mass distribution function remains a powerlaw
614: at all times, we may express $f(m,t) = c(t)m^{-q}$, such that
615: $m_L(t)$ is the growing upper limit of the distribution, $q$ is the slope
616: which is assumed to be constant (although see below), and
617: $c(t)$ is a normalization coefficient. Taking the lower limit of the mass
618: distribution to be $m_0$,
619: then moments as expressed in Eq. (2) are explicitly given for $q-p\neq 1$ by
620:
621: \begin{equation}
622: M_p(t) = \frac{c(t)}{1 + p -q}\left(m_L(t)^{1+p-q} - m_0^{1+p-q}\right).
623: \end{equation}
624:
625: \noindent
626: We then take the time derivative of Eq. (17) for the zeroth and second
627: moments, and substitute these expressions on the LHS of the corresponding
628: moment equations in Eq. (5) to get
629:
630: \begin{equation}
631: \frac{m_L^{1-q} - m_0^{1-q}}{1-q}\frac{dc}{dt} + cm_L^{-q}\frac{dm_L}{dt} =
632: -\frac{1}{2}c^2\Gamma_0(m_L),
633: \end{equation}
634:
635: \begin{equation}
636: \frac{m_L^{3-q} - m_0^{3-q}}{3-q}\frac{dc}{dt} + cm_L^{2-q}\frac{dm_L}{dt} =
637: c^2\Gamma_2(m_L),
638: \end{equation}
639:
640:
641: \noindent
642: where Eq. (18) is valid for $q\neq 1$, and $\Gamma_0$ and $\Gamma_2$ are the
643: integrals on the RHS of Eq. (5) for $k=0$ and $k=2$, respectively. That is
644:
645: \begin{equation}
646: \Gamma_0(m_L) = \int_{m_0}^{m_L(t)}\int_{m_0}^{m_L(t)}
647: K(m,m^{\prime})m^{-q}m^{\prime -q}\,dm\,dm^\prime,
648: \end{equation}
649:
650: \begin{equation}
651: \Gamma_2(m_L) = \int_{m_0}^{m_L(t)}\int_{m_0}^{m_L(t)}
652: K(m,m^{\prime})m^{1-q}m^{\prime 1-q}\,dm\,dm^\prime,
653: \end{equation}
654:
655: \noindent
656: where the kernel is given by
657: $K(m,m^\prime) = \sigma(m,m^\prime)\Delta V(m,m^\prime) S(m,m^\prime)$, with
658: $\sigma(m,m^\prime) = K_0(m^{1/3}+m^{\prime 1/3})^2$,
659: $K_0 = \pi(3/4\pi\rho_s)^{2/3}$, and $\rho_s$ is the particle material
660: density, assumed
661: constant. Note that for the special case of $q=1$, Eq. (18) has a slightly
662: different form which depends on $\ln{(m_L/m_0)}$.
663: After some simple algebra, Eq.'s (18) and (19) can be written as
664:
665: \begin{equation}
666: \frac{dm_L}{dt} = c\left[\frac{(3-q)(m_L^{1-q}-m_0^{1-q})\Gamma_2 +
667: \frac{1}{2}(1-q)(m_L^{2-q}-m_0^{2-q})\Gamma_0}{(3-q)(m_L^{1-q}-m_0^{1-q})
668: m_L^{2-q} - (1-q)(m_L^{3-q}-m_0^{3-q})m_L^{-q}}\right],
669: \end{equation}
670:
671: \begin{equation}
672: \frac{dc}{dt} = -c^2\left[\frac{(3-q)(1-q)(m_L^{-q}\Gamma_2 +
673: \frac{1}{2}m_L^{2-q}\Gamma_0)}{(3-q)(m_L^{1-q}-m_0^{1-q})
674: m_L^{2-q} - (1-q)(m_L^{3-q}-m_0^{3-q})m_L^{-q}}\right],
675: \end{equation}
676:
677: \noindent
678: which we integrate using a fourth order Runge-Kutta scheme. Equation (17) then
679: gives the moments $M_0$ and $M_2$ as a function of time, which can be
680: directly compared with direct integration of the same conditions
681: using the coagulation equation (Eq. 1).
682:
683: The advantage of this approach (in which a powerlaw is assumed at all
684: times) is its transparency; that is, the variables being sought
685: ($m_L$ and $c$) are solved
686: for directly. Furthermore, the change in the coagulation kernel as the
687: particle size distribution changes is included because the kernel is
688: updated and explicitly integrated into $\Gamma_0$ and $\Gamma_2$ with every
689: time step. This will prove
690: advantageous when additional effects such as sticking
691: are included in the kernel. In addition, source and sink terms need not
692: parameterized in terms of integer moments and can be implemented directly.
693: An unfortunate disadvantage of this approach is
694: that because the kernel must be integrated (in fact several times) over
695: both $m$ and $m^\prime$ every
696: time step to get $\Gamma_0$ and $\Gamma_2$, the CPU time involved is
697: significantly longer than a fully
698: implicit case (e.g., {\S} 2.1; also see {\S} 2.2.2); however, it
699: remains a much faster approach (orders of magnitude) than solving Eq. (1)
700: directly since the
701: cumbersome convolution has been eliminated.
702:
703:
704: \subsubsection{Approach 2: Semi-implicit Assumption}
705:
706: Here, we present an alternative moment-based approach to solving the realistic
707: coagulation kernel in which the integer moments appear directly.
708: Unlike the previous case of {\S} 2.2.1, where the double integral of
709: Eq. (5) is used to get the functions of $m_L$, $\Gamma_0$ and $\Gamma_2$,
710: we only integrate over one mass
711: variable (i.e., only integrate one of the integrals), defining the
712: functions
713:
714: \begin{equation}
715: C_0(m,t) = \int_{m_0}^{m_L} K(m,m^\prime)f(m^\prime,t) \,\,
716: dm^\prime,
717: \end{equation}
718:
719: \begin{equation}
720: C_2(m,t) = \int_{m_0}^{m_L} m^\prime K(m,m^\prime)f(m^\prime,t)
721: \,\,dm^\prime,
722: \end{equation}
723:
724: \noindent
725: where the form of the kernel $K(m,m^\prime)$ is the same that in {\S} 2.2.1.
726: We then fit $C_0$ and $C_2$ with a finite series in fractional
727: powers of $m$ in the same manner as given in Eq. (15). Substituting these
728: functions in place of one of the integrals (say over $m^\prime$), we may
729: integrate over $m$ to get
730:
731: \begin{eqnarray}
732: \frac{dM^\prime_0}{dt} = -\frac{1}{2M_0(0)}\int_{m_0}^{m_L(t)}
733: f(m,t)C_0(m,t)\,dm = -\frac{1}{2}\sum_i a_i \mu_{p_i}M^\prime_{p_i} \nonumber \\
734: \frac{dM^\prime_2}{dt} = \frac{1}{M_2(0)}\int_{m_0}^{m_L(t)} m f(m,t)C_2(m,t)
735: \,dm = \sum_i b_i \nu_{p_i+1}M^\prime_{p_i+1}, \nonumber \\
736: M^\prime_{p_i} = \left[M^\prime_0\right]^{\frac{1}{2}(p_i-1)
737: (p_i-2)}\left[M^\prime_1\right]^{-p_i(p_i-2)}\left[M^\prime_2\right]^
738: {\frac{1}{2}p_i(p_i-1)},
739: \end{eqnarray}
740:
741: \noindent
742: where we have made use of equations (8) and (9) to express the solution in
743: terms of the integer moments $k = 0,1,2$. Here, $\mu_{p_i} = M_{p_i}(0)/
744: M_0(0)$, $\nu_{p_i+1} = M_{p_i+1}(0)/M_2(0)$, $M^\prime_k = M_k(t)/M_k(0)$,
745: and $0\leq p_i \leq 1$. The $C_k$ are fairly smooth functions over a large
746: range of particle radii;, however, the accuracy of fitting a single series in
747: fractional moments over a very broad range of particle sizes
748: (i.e., over many orders of magnitude) may drop off significantly as the
749: broadness of the range increases. This issue may be circumvented by
750: employing a piecewise fit to the integrated kernels $C_k$.
751:
752: It is interesting to note that by this definition of $C_k$, we have
753: effectively accomplished what we set out to do in
754: our discussion at the beginning of {\S} 2.2, that is, defining the
755: coagulation kernel in terms of finite series in powers of the mass $m$.
756: The difference
757: here is that we have done so through the first integral of the kernel, and
758: not the kernel itself. This means that the mass distribution $f(m,t)$ is
759: expressed in the calulation of the $C_k$, but remains
760: implicit in the definition of the ODEs (Eq. 26). Thus, the method is
761: semi-implicit, because the RHS of the equations above can be
762: expressed in terms of the moments (as defined in Eq. 2). Equations (26)
763: are then integrated using the fourth order Runge-Kutta method, and may be
764: compared with the results of {\S} 2.2.1 and the direct integration of
765: Eq. (1).
766:
767: This semi-implicit approach tracks
768: the evolving kernel through the integration of Equations (24) and (25)
769: after every timestep, thus the computational time involved is similar
770: to the explicit case. In order to update the kernel,
771: one may solve for the new $m_L$ after each $\Delta t$
772: using the equation ($q\ne 1$)
773:
774: \begin{equation}
775: \frac{m_L-m_0}{m_L^{2-q}-m_0^{2-q}} - \frac{M_q}{(2-q)M_1} \simeq 0.
776: \end{equation}
777:
778: \noindent
779: The new
780: normalization coefficient $c$ can then be found from the definition of
781: $M_1 = \rho$.
782: We then reintegrate Equations (24) and (25) under the powerlaw assumption,
783: and then proceed to fit the $C_k(m,t)$ with a finite series in fractional
784: powers of $m$. Although, in principle, any other two moments could be
785: used to obtain $m_L$, $M_q$, which lies between $M_2$ and $M_1$, and because
786: it roughly characterizes the evolution of the largest particle (see discussion
787: at the end of {\S} 2.2), seems the most
788: consistent choice. The $q$-th moment is calculated using the Lagrange
789: polynomial interpolation scheme (Eqs. 8 and 9).
790:
791: The advantage of the moments method lies in the ability to express the
792: differential equations in terms of the moments of the
793: distribution (i.e., their integrated properties).
794: If a more explicitly
795: mass-dependent approach is adopted (as is the case in {\S} 2.2.1, and the
796: semi-implicit approach described here),
797: then the computational time significantly increases.
798: One can improve the speed of computation by calculating the $C_k$
799: periodically, or in the extreme case, only at $t=0$ which would make
800: the approach truly {\it implicit} (e.g., {\S} 2.1). The advantage of an
801: implicit approach
802: is that it becomes fully general
803: (the form of $f$ is only assumed at the onset), and
804: also in the time it takes to solve ($< 1$ minute). The bulk of the time is
805: spent in the
806: integration of equations (24) and (25) which would occur only once.
807: The disadvantage, of course, is that the particle velocity distribution is
808: not updated as it changes with time (due to, e.g., changes in the bounds
809: of the size distribution). We present examples of both extremes in {\S} 3.
810:
811: If one wanted to implement a mass- or velocity-dependent sticking
812: coefficient $S$, it can readily be included in the integration to
813: obtain the $C_k$. The additional inclusion of source and sink terms
814: due to erosion, fragmention, or
815: gravitational growth in this semi-implicit formalism would require that we fit
816: these terms in a similar manner to the $C_k$ so that their subsequent
817: integration over all $m$ will yield sums over integer moments weighted by
818: different sets of coefficients. Caution must be exercised in fitting, e.g.,
819: the gravitational growth term to ensure
820: that the system of equations remains closed. Under such circumstances, a
821: fully explicit approach such as that of {\S} 2.2.1 may be preferred.
822: Alternatively, these effects may be included in particle-histogram space
823: in between coagulation interations.
824:
825: Finally,
826: we should point out that allowing for other parameters (such as the index
827: of the powerlaw $q$)
828: to vary with time, does not pose a problem in either of
829: the approaches we have presented. In both cases, one would simply need an
830: additional moment, e.g. $M_3$, to determine $q(t)$. Similarly, a bifurcated
831: distribution in which the powerlaw exponent changes at a particular particle
832: size \citep[see, e.g.,][]{ken99} may also be studied.
833:
834: In {\S} 3, we will compare the
835: two approaches to the direct integration of the coagulation equation for
836: cases in which there are
837: only systematic velocities ($v_T = 0$), as well as cases in which the
838: velocity differences are induced by turbulent motions.
839:
840:
841: \section{Numerical Results}
842:
843: We carried out several calculations in order to demonstrate the accuracy
844: of each alternative method compared to the brute force integration of the
845: collisional coagulation equation. For the purposes of comparison,
846: unless otherwise noted, we chose the initial conditions to be a minimum mass
847: solar nebula at 1 AU at a height of $z=10^3$ km above the midplane, and a
848: particle size distribution with a minimum initial radius of 1 cm and
849: a maximum initial radius of 10 cm. The
850: powerlaw exponent $q = 11/6$, which is assumed to be constant in these
851: calculations, is representative of a fragmentation population.
852: Standard integrations were carried out with a timestep of $\Delta t = 10$
853: years.
854:
855:
856: \subsection{Laminar Case ($v_T = 0$)}
857:
858: In Figure 2, we compare the case in which there are only systematic
859: (pressure-gradient driven) velocities
860: between particles ($v_T=0$). The solid curves represent the explicit
861: assumption (invariant powerlaw slope $q$ assumed in integration of
862: Equations 22 and 23), while the dashed curves represent the
863: implicit assumption. That is, in the integration of Eq. 26, the form of
864: the mass distribution $f$ is assumed only at $t=0$. As before, the symbols
865: represent the brute force calculation for two different grid resolutions
866: (solid = 100 bins, open = 1000 bins). Since with the explicit assumption we do
867: not solve for the moments specifically, the values for the solid curves were
868: obtained by substituting the time integrated values of $c(t)$ and $m_L(t)$
869: back into Eq. (17) for $k = 0,2$ only. This is because
870: although we have used the moment equations to obtain these results, the
871: explicit assumption of an invariant powerlaw size distribution means that
872: only the equations for the moments $M_0, M_2$ are needed (although see
873: {\S} 3.2). This is not true, however, for the implicit (and semi-implicit)
874: assumption, where
875: all the moments $M_0,M_1,M_2$ appear in the differential equations.
876:
877: We see that there is excellent agreement between both approaches and the
878: numerical values obtained with the highest resolution case. In fact, the
879: agreement between the explicit and implicit assumptions is also
880: quite good. However, at this stage there has not been a great deal of
881: growth (the largest particle size in the distribution has only grown
882: to $r_L = 11$ cm in size by the end of the simulation for the explicit case).
883: Note that even though
884: in the implicit case the $C_k$ are calculated only once, the estimate for the
885: largest mass $m_L$ for the evolving distribution $f(m,t)$ is still calculated
886: using Eq. (27). At least in the case of minimal or slow growth, it appears
887: that an implicit approach, or even periodic calculation of the $C_k$, may be
888: sufficient.
889:
890: Numerical glitches in the low resolution brute force case arise from the
891: interpolation
892: scheme for sampling the kernel. In particular, these glicthes are
893: likely enhanced because the systematic
894: relative velocities quickly approach zero
895: for identical particle sizes (with the effect much more prominent for larger
896: particles, hence not appearing so much in $M_0$). The higher resolution case
897: contains enough points to smooth out this effect.
898:
899: Note that (in all simulations) the direct integration of the coagulation
900: equation gives a constant $M_1$
901: as is to be expected given that we found $dM_1/dt = 0$ in the
902: derivation of the moment equations in the absence of sources and sinks; this
903: further validates the numerics
904: of the brute force solution even
905: for the complicated collisional kernel being utilized, and also provides
906: a posteriori validation of our result that $M_1=\rm{constant}$ from,
907: e.g., Eq. (5), which further validates derivation of equations (24)
908: and (25) which was based on symmetry of the kernel.
909:
910:
911: \subsection{Turbulence Case ($v_T \neq 0$)}
912:
913: Next, we explored cases in which the systematic velocities were set to zero
914: so that velocity differences between particles are due only to those induced
915: by turbulence. Depending on the magnitude of the turbulence parameter
916: $\alpha$, these induced velocities can be either large or small relative to the
917: systematic velocities. To demonstrate, we ran cases for three
918: different values $\alpha = 10^{-6},10^{-5}$, and $10^{-4}$. In the absence
919: of any mechanism to counter growth (e.g., fragmentation), larger $\alpha$
920: translates to faster rates of growth (due to larger relative
921: velocities).
922:
923: In Figure 3 we plot the results for $\alpha = 10^{-4}$,
924: for the explicit (solid curves) and implicit (short dashed curves)
925: approaches. The growth rate is more rapid than in the laminar case, with
926: the second scaled moment $\sim 70$\% larger (compared to Fig. 2) at the
927: end of the run, meaning that the
928: largest particle size achieved is $r_L \sim 13.5$ cm. We note
929: that, as was the case for the $M^\prime_2$ curves in Fig. 2, the explicit
930: and implicit cases are very similar; however, this is not the case
931: for the $M^\prime_0$ curves. The explicit approach overestimates the the
932: value of $M^\prime_0$ compared to the highest resolution brute force
933: calculation, whereas the implicit approach significantly underestimates
934: the zeroth moment. Both approaches understimate the value of $M^\prime_2$.
935: When we compare integrations of $M^\prime_2$ for smaller
936: values of $\alpha$ as we do in Figure 4, we find that both approaches fit
937: the coagulation calculation quite well. The growth rate for these two
938: values of $\alpha$ are more in line with growth rate found when there are
939: only systematic velocities, so the agreement should not be surprising.
940:
941: The long-dashed curves in Figures 3 and 4
942: are the implementation of the semi-implicit approach ({\S} 2.2.2) in which
943: the $C_k$ are updated at every time step. In this case, the semi-implicit
944: approach provides a much a much better fit to the brute force calculation,
945: indicating that the kernel is evolving fast enough that an implicit
946: approach cannot capture this effect.
947: We consider the slight discrepencies between the
948: semi-implicit approach and the brute force calculation to be as much a
949: result of grid resolution as inaccuracy in using the Lagrange polynomial
950: fits to the fractional moments. The explicit and fully implicit approach
951: values of $r_L$ apparently understimate the largest particle size
952: relative to the semi-implicit approach (which gives a value of
953: $r_L \simeq 14$ cm).
954:
955:
956: Finally, we explored a variation of the explicit approach in which
957: the condition $M_1 = \rho$ was strictly enforced (recall that
958: $M_1$ does not appear in Equations 18-19). This amounts to replacing
959: Eq. (18) (derived for $M_0$) with the corresponding equation for $M_1$.
960: As a result, we cannot simultaneously fit both $M_2$ and $M_0$.
961: Alternatively, we can use only Eq. (19) for the second moment, by substituting
962: $c = (2-q)\rho/(m_L^{2-q}-m_0^{2-q})$ (from the definition of $M_1$, Eq. 17)
963: in Eq. (19) so that Eq. (19) alone
964: determines the growth of the largest particle. The differential
965: equation for $m_L$ then becomes
966:
967: \begin{equation}
968: \frac{dm_L}{dt} = \left[\frac{(3-q)(2-q)\rho\Gamma_2}
969: {(3-q)(m_L^{2-q}-m_0^{2-q})m_L^{2-q} - (2-q)(m_L^{3-q}-m_0^{3-q})
970: m_L^{1-q}}\right].
971: \end{equation}
972:
973: \noindent
974: The results of the integration of Eq. (28) are shown as the dotted curves
975: in Figure 3 and for the $\alpha = 10^{-4}$ case in Figure 4. It is clear that
976: this modification provides a much better fit to $M_2$, perhaps even better
977: than the semi-implicit case above. However, using
978: $m_L(t)$ calculated from Eq. (28), and solving for $c(t)$ does a poor job
979: fitting $M_0$ (see Figure 3). Thus we conclude that although the
980: modification of the explicit approach matches the brute force calculation of
981: $M_2$ quite well, only the semi-implicit case is able to provide a
982: simultaneous fit to both the zeroth and second moments (under the
983: condition that $M_1=\rho = {\rm{constant}}$).
984:
985:
986: \subsection{A Model Comparison}
987:
988: \citet{gar07} has developed a simplified analytical approach for dealing
989: with the
990: growth of particles in a turbulent regime, in which the particle size
991: distribution is parameterized by a powerlaw in particle mass with the same
992: exponent we have used in this paper ($q = 11/6$).
993: Similar to what we have presented in previous sections, the underlying
994: assumption of \citet{gar07} is that collisions between particles occur
995: frequently enough that a steady-state balance is reached in the form of
996: the particle
997: size distribution, but with an upper size cutoff that varies with time.
998: Thus, the model of \citet{gar07} is {\it explicit} and follows only two
999: parameters: the growth
1000: of the largest particle $m_L(t)$, and the normalization factor $c(m_L,t)$.
1001:
1002: Given the similarity of the underlying assumptions for the Garaud model
1003: and the examples we have presented, it is a
1004: useful exercise to compare the results of the two approaches directly.
1005: The growth of the largest particle $r_L$ in the Garaud model (her Eq. 36),
1006: expressed in terms of the notation used in this paper, is given by
1007:
1008: \begin{equation}
1009: \frac{dr_L}{dt} = \frac{\rho \Omega H}{\rho_s} \sqrt{\frac{\alpha
1010: \gamma {\rm{St}}_L}{1 + 64{\rm{St}}^2_L(2 + 5{\rm{St}}_L^{-0.1})^{-2}}}.
1011: \end{equation}
1012:
1013: \noindent
1014: In Eq. (29) ${\rm{St}}_L$ is the Stokes number for the largest particle
1015: $r_L = (3m_L/4\pi\rho_s)^{1/3}$, $H=c_g/\Omega$ is the scale height of the gas,
1016: $\gamma$ is the adiabatic index of the gas, and
1017: in this expression it is assumed that the sticking coefficient $S = 1$, and
1018: that $m_L>>m_0$. The above
1019: equation is similar to the formula for grain growth proposed by \citet{ste97}
1020: to factors of order unity. Finally, we point out that
1021: the Garaud model for particle growth
1022: is restricted to the value $q = 11/6$ in order to preserve its completely
1023: analytical nature.
1024: As a means of a fair comparison, we chose to compare the Garaud model to
1025: our explicit case (Eq. 28) for reasons explained below. In the limit of
1026: $m_L >> m_0$ and $q=11/6$, Eq. (28) becomes
1027:
1028: \begin{equation}
1029: \frac{dr_L}{dt} \simeq 0.05 \frac{\rho}{\rho_s}\frac{\Gamma_2}{m_L}.
1030: \end{equation}
1031:
1032: We present
1033: the results of our comparison in Figure 5 for the same initial conditions
1034: as described at the beginning of {\S} 3.2 (upper curves).
1035: We find quite generally that our explicit calculation (Eqs. 28 and 30)
1036: leads to a faster growth rate than what is predicted from the Garaud model,
1037: initially. We note that
1038: the Garaud expression steepens quickly for $r_L \gtrsim 13$ cm,
1039: suggesting that the growth rates of the two approaches are more comparable at
1040: later times.
1041: However, the minor ``kink'' in
1042: the long-dashed curve is due to the shift from Epstein to Stokes flow. A
1043: much more subdued kink is visible in the explicit approach
1044: which uses the full expressions for the turbulent velocity, whereas in the
1045: derivation of Eq. (29), it is
1046: assumed that the stopping time in the Stokes regime is defined
1047: by some mean characteristic velocity which leads to a much more
1048: noticeable discontinuity.
1049: Regardless, the overall more subdued growth rate elicited by Eq. (29) (despite
1050: the steepening at later times due to a shift in flow regimes) is apparent
1051: from the lower set of curves in Fig. 5 where
1052: the initial conditions were chosen with $r_L(0) = 1$ cm. For this case, growth
1053: occurs only in the Epstein regime, but still, the curves for the explicit
1054: approach
1055: and that calculated from Eq. (29) begin to diverge. Thus,
1056: the Garaud expression apparently underestimates the growth rate relative to
1057: our approach.
1058:
1059: We emphasize that the treatment of particle growth by \citet{gar07} requires
1060: several approximations in order to derive a purely analytical expressions for
1061: $dr_L/dt$. Besides the aforementioned restriction of $q=11/6$, Garaud
1062: approximates the full expressions for the turbulent velocities that we
1063: use here by partitioning $v_T$ into seperate cases dependent on
1064: the particles' stopping time relative to the turnover times of the smallest
1065: and largest scale eddies. Furthermore, some question may be raised as to the
1066: comparability of the moment equations used here to derive Eq. (28), versus
1067: the particle growth equation used by \citet[her Eq. 29]{gar07}.
1068: The equation used by \citet{gar07}
1069: is more akin to a ``sweep-up'' equation, with no sources or sinks, than to a
1070: formal coagulation equation. Because
1071: it bears some resemblance to the equation for $M_2$ (with some algebra,
1072: to factors of order unity), we concluded that our Eq. (28) is the appropriate
1073: analog. Despite the differences in growth rate, we find the agreement in
1074: the general trend of growth of the two approaches reassuring.
1075:
1076:
1077:
1078: \section{Opacity Calculations}
1079:
1080: Evolutionary models of protoplanetary nebulae, giant planet atmospheres, etc.
1081: must somehow treat the escape of thermal radiation \citep{pol96,hub05,dur07}.
1082: In most cases, particles provide the primary opacity for these
1083: models. Observations of these and similar objects often rely on Spectral Energy
1084: Distributions (SEDs) which can be compared to a model once the model's internal
1085: temperature distribution is known; clear evidence is seen for grain growth in
1086: many cases \citep[see review by][]{nat07}. Because of the nearly
1087: insurmountable
1088: computational burden involved with performing a fully self-consistent
1089: calculation of particle growth by coagulation along with an already difficult
1090: fluid dynamical calculation, most modelers simply assume some invariant
1091: particle size distribution, such as the MRN interstellar grain distribution, or
1092: make arbitrary assumptions about particle growth \citep{hub05}.
1093:
1094: In the simplest regime (monodisperse particle radius $r$ larger than a
1095: wavelength), the particle opacity can be written as the area per unit
1096: mass:
1097:
1098: \begin{equation}
1099: \kappa = \frac{3}{4 \rho_s r} \hspace{0.1 in} {\rm cm}^2\,{\rm g}^{-1};
1100: \end{equation}
1101:
1102: \noindent
1103: thus growth in radius from 0.1$\mu$ to 1 mm leads to a factor of $10^4$ change
1104: in opacity. To the degree that this wavelength-independent regime holds,
1105: including particle size evolution by the moments method in one's
1106: evolutionary models would allow a very simple way to track particle growth and
1107: decreasing opacity. For instance, equation (31) above is easily generalized to
1108: the area per unit mass integrated over the size distribution:
1109:
1110: \begin{equation}
1111: \kappa = \frac{\int \pi r^2 f(m)\, dm}{\int m f(m)\, dm}
1112: = \left(\frac{9 \pi}{16 \rho_s^2}\right)^{1/3}\frac{M_{2/3}}{M_1}.
1113: \end{equation}
1114:
1115: \noindent
1116: As an application of the moments method in Figure 6, we have calculated
1117: the decrease
1118: in opacity (given by Eq. 32) with time using the semi-implicit approach.
1119: An initial particle size distribution with a lower bound of $r_0=0.1$ cm
1120: and $q=11/6$ was used. Both the pressure gradient driven systematic
1121: velocities and the
1122: turbulence-induced velocities were used. In the absence of any mechanism to
1123: hinder particle growth, larger values of $\alpha$ lead to more steeply
1124: decreasing opacities with time.
1125:
1126:
1127: In a regime where the particle extinction efficiency $Q(r,\lambda)$ is
1128: wavelength-dependent (say, if the particles are comparable to or smaller than
1129: the wavelength), one simply integrates $Q(r,\lambda)$ over the powerlaw mass
1130: distributions resulting from the moments model. For example,
1131: \begin{eqnarray}
1132: \kappa_{\lambda} = \frac{\int_{m_0}^{m_L} \pi r^2 Q(r,\lambda) f(m)\, dm}
1133: {\int_0^{m_L} m f(m)\, dm}
1134: = \frac{1}{M_1}\int_{m_0}^{m_L} \pi r^2 Q(r,\lambda) c(m_L)m^{-q}\, dm
1135: \nonumber \\
1136: = \frac{c(m_L)}{M_1}\left(\frac{9 \pi}{16 \rho_s^2}\right)^{1/3}
1137: \int_{m_0}^{m_L} Q(r,\lambda)m^{2/3 -q } dm.
1138: \end{eqnarray}
1139:
1140: \noindent
1141: These opacities $\kappa_{\lambda}$ can be used to calculate Planck or Rosseland
1142: (wavelength-averaged) means for use in radiative transfer models.
1143: Recall that the powerlaw slope
1144: $q$ can be freely adjusted within a small but plausible range to explore
1145: different growth regimes.
1146:
1147: \section{Porosity}
1148:
1149: Fractal growth of particles by low-velocity sticking of small solid monomers
1150: with radius $r_o$ and mass $m_o$ (and/or aggregates of such monomers) causes
1151: them to have a density much less than the material density of the monomers
1152: \citep{bec00,dom07,orm08}. These porous
1153: particles can be described as fractals with dimension $D$, such that the
1154: particle mass $m$ increases proportionally to $r^D$ where $r$ is some effective
1155: radius and $D$ is the fractal dimension. Thus the particle's internal density
1156: is a function of particle size:
1157:
1158: \begin{equation}
1159: \rho(r) = \frac{3 m }{4 \pi r^3} \sim \frac{3 m_o (r/r_o)^D}
1160: {4 \pi r^3} = \frac{3 m_o}{4 \pi r_o^3}(r/r_o)^{D-3} = \frac{3 \rho_o}
1161: {4 \pi}(r/r_o)^{D-3}.
1162: \end{equation}
1163:
1164: \noindent
1165: For a typical situation where $D \sim 2$ \citep{dom07},
1166: $\rho(r) \propto r^{-1}$ and thus the
1167: product $r \rho(r)$ is a constant across a wide range of particle sizes (until
1168: compaction sets in). These more complex but quite plausible particle
1169: density-size relationships complicate the expressions for particle stopping
1170: time and Stokes number ({\S} 2.2, Eq. 11; also, see appendix).
1171:
1172: Because the particle stopping times enter in through the collisional kernel,
1173: the fractal nature of particles can be accounted for using the method of
1174: moments in a straightforward manner while maintaining the criterion for
1175: closure of the system. We can verify this by noting that the mutual particle
1176: cross section
1177:
1178: \begin{equation}
1179: \sigma (m,m^\prime) = \frac{\pi r_o^2}{m_o^{2/D}}\left(m^{1/D} +
1180: m^{\prime 1/D}\right)^2,
1181: \end{equation}
1182:
1183: \noindent
1184: is proportional to $m^{2/D}$, whereas the stopping time (Eq. 11) in the Epstein
1185: regime (the regime that would apply to fluffy fractal aggregates) where
1186: the drag force $F_D = (4/3)\pi r^2 c_g \rho_g \Delta V_{pg}$
1187: \citep[e.g.,][]{cuz93} is given by
1188:
1189: \begin{equation}
1190: t_s = \frac{3m}{4\pi r^2 c_g\rho_g} = \frac{3m_o^{2/D}}{4\pi r_o^2 c_g\rho_g}
1191: m^{1-2/D}.
1192: \end{equation}
1193:
1194: \noindent
1195: In the limit of small Stokes number
1196: ($\rm{St}\ll 1$), both the systematic and turbulence-induced velocities
1197: are proportional to $\rm{St}$, so that $\Delta V \propto t_s \propto
1198: m^{1-2/D}$, and the entire kernel is proportional to $m$. For
1199: larger $\rm{St}$, the kernel has a shallower powerlaw
1200: dependence. Thus, in general, the dependence of $K(m,m^\prime)$ on the
1201: mass is $\lesssim m$, preserving closure of the system, even for fractal
1202: particles. Even though the variation in the properties of the evolving
1203: distribution due to porous particles are incorporated directly into the
1204: kernel, and are folded into the integration of the explicit approach, the
1205: effects of fractal aggregates can, nonetheless, affect
1206: which moments characterize what properties in the semi-implicit (or implicit)
1207: approach. For
1208: example, the wavelength-independent opacity expressed in Eq. (32) takes
1209: the form $\kappa = (\pi r_o^2/m_o^{2/D})M_{2/D}/M_1$.
1210:
1211:
1212: It should be noted that the value $D=2$ represents a special case in that
1213: the particle-to-particle relative velocities in the Epstein regime do not
1214: depend on the mass of the fractal particle. Indeed, this would seem to
1215: indicate that fractal growth can proceed unabated with the corresponding
1216: stopping time of the fluffy aggregate remaining the same as that of a single
1217: monomer, which would have a significant effect on other particle properties. In
1218: particular, the wavelength-independent opacity for $D=2$ is constant. However,
1219: impacts will eventually lead to compaction or even fragmentation depending
1220: on the relative velocities \citep[e.g.,][]{orm07b}.
1221: Both fractal grains and non-fractal particles in the same mass distribution
1222: can be treated in the explicit approach without any modifications, while a
1223: piecewise fit to
1224: the integrated kernel $C_k(m)$ ({\S} 2.2.2) can be used in order to
1225: account for the change in regimes in the semi-implicit approach.
1226:
1227:
1228:
1229: \section{Conclusions}
1230:
1231:
1232: We have demonstrated an approach to solving the collisional coagulation
1233: equation with an arbitrary collisional kernel which should be useful in cases
1234: when it is only necessary to keep track of general
1235: properties of the distribution. This approach involves solving a finite set of
1236: coupled differential equations in terms of the integer moments of the particle
1237: size distribution. The number of equations (and thus moments) needed depends
1238: on the number of properties being tracked. The advantage of the moments method
1239: approach is that it allows for considerable savings in computational
1240: time compared to direct integration of the coagulation equation,
1241: which requires keeping track of every particle size at every spatial location
1242: and timestep.
1243:
1244: In this paper we have specifically
1245: studied the growth of the largest particle under the assumption that
1246: the particle size distribution is a powerlaw; however, the technique can be
1247: extended to track other properties of the distribution that may change with
1248: time. There are many reasons to believe that a powerlaw size distribution is
1249: a natural end-state of particle growth, especially those with equal mass
1250: per decade, because they have self-preserving properties \citep{cuz06}.
1251: With the assumption of a powerlaw distribution, we
1252: have provided two different approaches to solving the moment equations,
1253: one explicit in which the powerlaw assumption is enforced rigorously at
1254: all times, and a semi-implicit approach in which the kernel is integrated
1255: over one of the mass variables as much as once every time step. The latter
1256: approach can be made fully implicit by only assuming the form of the mass
1257: distribution at $t=0$. These approaches
1258: are significantly faster than solutions of the coagulation equation
1259: because, in particular, the convolution integral (first term on RHS of Eq. 1)
1260: has been eliminated. In realistic evolutionary models, intermediate steps
1261: performed in particle ``histogram'' or ``size-distribution'' space may be
1262: interleaved with moments-based coagulation steps in order to account for,
1263: e.g., advective/transport terms.
1264:
1265: We have compared
1266: these alternate approaches to the brute force integration of the full
1267: coagulation equation for cases in
1268: which there are only systematic velocities, and cases in which the differences
1269: in velocity between particles is induced by turbulence. If the growth rate
1270: is gradual, the explicit and implicit
1271: approaches match the brute force calculation well (Fig. 2),
1272: whereas
1273: faster growth rates are more difficult to model (Fig. 3). We find that we
1274: are able to use the semi-implicit approach in which the $C_k$ are updated
1275: at every timestep, and a modification to the explicit approach, in which
1276: we solve the equation for $M_2$ only with the assumption
1277: of $\rho=M_1$ strictly enforced,
1278: in order to compensate for the faster growth rates.
1279: The modification to the explicit approach is useful if
1280: one is not particularly interested in following the evolution of the number
1281: density of particles, or other properties which may be decribed by
1282: (fractional) moments $< 1$ (e.g., see {\S} 4). Our results also suggest
1283: that a fully implicit approach is probably most useful under circumstances in
1284: which the kernel depends on the mass in a straightforward manner (e.g, the
1285: Saffman-Turner kernel, {\S} 2.1).
1286:
1287: We have compared the approaches developed in this paper to an alternative
1288: model for particle growth in a turbulent nebula
1289: \citep{gar07}. We have found that there is fairly good agreement in general,
1290: but that the curves diverge as time proceeds. This does not appear to be
1291: particle-size dependent, or due to a shift in the flow regime.
1292: The Garaud expression (Eq. 29) underestimates the growth rate of
1293: the largest particle size relative to our approach by only $\sim 20-30$\%
1294: in our comparison. We note that the advantage
1295: of the method of \citet{gar07} is
1296: that it is purely analytical; however, preservation of her analytical
1297: approach requires, amongst other things, the powerlaw
1298: exponent be restricted to $q=11/6$. Our approach has no such restriction
1299: on the choice of exponent $q$, nor for $q$ to even be a constant.
1300:
1301: As a sample application, we show how the moments method can
1302: be used in one's evolutionary model to track
1303: particle growth and opacity. As a specific case, we calculated
1304: the change in wavelength-independent opacity with time for an initial
1305: particle size distribution with upper and lower bounds of $0.1-1$ cm.
1306: Both systematic and turbulent velocities were included. In the absence of
1307: any mechanism to counter growth, the opacity decreases sharply for higher
1308: choices of the turbulent parameter $\alpha$. Extension to cases in
1309: which the extinction efficiency is wavelength dependent is straightforward.
1310: Such opacities can be used to calculate Planck or Rosseland
1311: wavelength-averaged means for use in radiative transfer codes.
1312:
1313: Finally, we indicate how porous particles with fractal dimension $D$ can
1314: be accounted for in the moments method. The particle-size density
1315: relationships that arise affect the particle stopping times and Stokes
1316: number, both of which appear only in the collisional kernel. Thus
1317: implementation is straightforward. We can treat mass distributions composed
1318: of only fractal grains, or both fractal and non-fractal particles.
1319: In either case, relatively little modification is needed in the explicit
1320: approach, whereas with the semi-implicit (and implicit) approach, a piecewise
1321: fit to the integrated kernel $C_k$ using the method as described in {\S} 2.2.2
1322: can be used to account for the change in particle growth regime when both
1323: types of particles are included.
1324:
1325: The computational burden of directly solving the coagulation equation
1326: makes it
1327: quite prohibitive to explore large regions of parameter space, and thus
1328: it serves
1329: as the primary bottleneck in evolutionary growth models. The approach
1330: demonstrated herein is intended to obtain robust, quantitative results for
1331: disk properties such as particle growth
1332: timescales and ``typical'' particle sizes that may be used in modeling
1333: efforts that are
1334: focused more on the larger problem of planetesimal formation.
1335:
1336: Although we have not included sticking or fragmentation in this paper, we
1337: believe that the moment equations (Eq. 5) will remain valid at least up
1338: to the fragmentation barrier. This size will depend on the
1339: assumed particle strengths and choice of nebula parameters. The treatment
1340: of growth up to the fragmentation size is consistent with recent work by
1341: \citet{bra08} for the case in which additional effects such as radial
1342: drift are not included. The model as presented in this paper, however,
1343: should be useful for the early stages of protoplanetary nebula particle
1344: growth relevant to spectral energy distributions and MRI suppression.
1345:
1346: In a forthcoming paper we will explore the effects of a
1347: variable sticking coefficient $S(m,m^\prime)$. Furthermore, we will explore the
1348: addition of source and sink terms such as gravitational growth,
1349: erosion, sublimation, and condensation in addition to fragmentation. The
1350: ultimate goal will be to apply this methodology to
1351: a global model that studies the evolution of both the gas and solids in
1352: nebular and subnebular (giant planetary) environments.
1353:
1354:
1355: \acknowledgements{We wish to thank Sandy Davis, Fred Ciesla, and Olenka
1356: Hubickyj for internal reviews which improved the exposition of this paper,
1357: and an anonymous reviewer for his or her careful analysis of the manuscript.
1358: This work was supported by a grant from NASA's
1359: Origins of Planetary Systems Program.}
1360:
1361: \appendix
1362:
1363: \section{Generalization of the Systematic (Pressure Gradient
1364: Driven) Velocities}
1365:
1366: In this appendix, we generalize the basic equations of \citet[also, see
1367: Tanaka et al. 2005]{nak86} for a two-component fluid by extending
1368: the particle component to incorporate a size distribution.
1369: We adopt cylindrical coordinates $(R,\phi,Z)$, designating the corresponding
1370: gas velocity components as
1371: $(u,v,w)$ and the particle velocity components as $(U_i,V_i,W_i)$.
1372: The equations of motion of
1373: particles and gas for the radial and tangential velocities, generalized for
1374: a particle size distribution, are given by
1375:
1376: $$
1377: \frac{\partial U_i}{\partial t} = -A_i \rho_g(U_i - u) + 2\Omega V_i,
1378: \eqno{(\rm{A}1)}
1379: $$
1380: $$
1381: \frac{\partial V_i}{\partial t} = -A_i \rho_g(V_i - v) - \frac{1}{2}\Omega
1382: U_i,
1383: \eqno{(\rm{A}2)}
1384: $$
1385: $$
1386: \frac{\partial u}{\partial t} = -\sum_j A_j \rho_{j}(u - U_j) + 2\Omega v -
1387: \frac{1}{\rho_g}\frac{\partial p_g}{\partial R},
1388: \eqno{(\rm{A}3)}
1389: $$
1390: $$
1391: \frac{\partial v}{\partial t} = -\sum_j A_j \rho_{j}(v - V_j) - \frac{1}{2}
1392: \Omega u,
1393: \eqno{(\rm{A}4)}
1394: $$
1395:
1396: \noindent
1397: where the system is assumed to be axysimmetric. Here, $A_i =
1398: (\rho_gt_{{\rm{s}}i})^{-1}$ with $t_{{\rm{s}}i}$ the stopping time of a
1399: particle of radius
1400: $r_{i}$, $\rho_g$ and $p_g$ are the gas mass density and pressure,
1401: and $\rho_{j}$
1402: is the material density of particles of radius $r_{j}$. We have not included
1403: the
1404: vertical component equations; here, generalization to a particle size
1405: distribution
1406: is straightforward since it remains true in general that $|w/W_i| << 1$; that
1407: is, the vertical gas component of the velocity $w$ is negligibly small
1408: compared with $W_i$ \citep[see appendix,][]{nak86}.
1409:
1410: If the dust stopping time is short compared to the orbital period, we can seek
1411: steady-state solutions for the velocity components. Setting $\partial/\partial
1412: t = 0$, this system can be solved exactly in the Epstein regime. By solving
1413: Eq. (A1) for $V_i$ and inserting into Eq. (A2), we obtain
1414:
1415: $$
1416: U_i = \frac{A_i^2\rho_g^2u + 2\Omega A_i\rho_gv}{A_i^2\rho_g^2 + \Omega^2} =
1417: \frac{u + 2v{\rm{St}}_i}{1 + {\rm{St}}_i^2},
1418: \eqno{(\rm{A}5)}
1419: $$
1420:
1421: $$
1422: V_i = \frac{-(\Omega/2)A_i\rho_gu + A_i^2\rho_g^2v}{A_i^2\rho_g^2 + \Omega^2}
1423: = \frac{v - (1/2)u{\rm{St}}_i}{1 + {\rm{St}}_i^2},
1424: \eqno{(\rm{A}6)}
1425: $$
1426:
1427: \noindent
1428: where $V_i$ was obtained by insertion of $U_i$ back into Eq. (A1). In
1429: Eqs. (A5)-(A6) we have expressed the last equality in terms of the Stokes
1430: number ${\rm{St}}_i = t_{si}\Omega$, with $\Omega$ being the local Kepler
1431: frequency. These expressions for the
1432: individual particle velocity components can be inserted into Eqs. (A3) and
1433: (A4) to yield expressions for radial and tangential gas velocity components.
1434: With little difficulty, one finds
1435:
1436: $$
1437: u = 2\eta v_K \frac{s_1}{s_1^2 + (1 + s_0)^2},
1438: \eqno{(\rm{A}7)}
1439: $$
1440:
1441: $$
1442: v = -\eta v_K \frac{1 + s_0}{s_1^2 + (1 + s_0)^2},
1443: \eqno{(\rm{A}8)}
1444: $$
1445:
1446: \noindent
1447: where
1448:
1449: $$
1450: s_0 = \sum_j \frac{A_j^2\rho_g\rho_{j}}{A_j^2\rho_g^2 + \Omega^2} =
1451: \sum_j \frac{\rho_{j}}{\rho_g}\frac{1}{1 + {\rm{St}}_j^2},
1452: \eqno{(\rm{A}9)}
1453: $$
1454:
1455: $$
1456: s_1 = \Omega \sum_j \frac{A_j\rho_{j}}{A_j^2\rho_g^2 + \Omega^2} =
1457: \sum_j \frac{\rho_{j}}{\rho_g}\frac{{\rm{St}}_j}{1 + {\rm{St}}_j^2}.
1458: \eqno{(\rm{A}10)}
1459: $$
1460:
1461: \noindent
1462: and $\eta = -(1/2\rho_g\Omega^2R)(\partial p_g/\partial R)$. These expressions
1463: agree with those obtained by \citet{tan05} with the exception that
1464: their expressions for $s_0$ and $s_1$ are expressed in integral form.
1465:
1466: From these expressions, one can easily obtain expressions for the relative
1467: velocities with respect to the gas $U_i-u$ and $V_i-v$, or the relative
1468: velocities between particles of different sizes $U_i-U_j$ and $V_i-V_j$.
1469: These expressions are also applicable for the Stokes regime in which the
1470: stopping times are a function of the relative particle-gas velocities
1471: (cf. Eq. 11). In
1472: this case, iterations must be performed in order to obtain the correct
1473: relative velocities. Convergence is generally achieved in a small number
1474: of iterations.
1475:
1476:
1477: \begin{thebibliography}{}
1478:
1479: \bibitem[Beckwith \& Sargent(1991)]{bec91} Beckwith, S. V. W., \& Sargent,
1480: A. I. 1991, \apj, 381, 250
1481: \bibitem[Beckwith et al.(2000)]{bec00}
1482: Beckwith, S. V. W., Henning, T., \& Nakagawa, Y., In Protostars and Planets
1483: IV, V. Mannings, A. P. Boss, and S. S. Russell, eds.,
1484: (Tucson: Univ. of Arizona Press), 533
1485: \bibitem[Brauer et al.(2008)]{bra08}
1486: Brauer, F., Dullemond, C. P., \& Henning, Th., 2008, \aap, 480, 859
1487: \bibitem[Bromley \& Kenyon(2006)]{bro06}
1488: Bromley, B. C., \& Kenyon, S. J., 2006, \aj, 131, 2737
1489: \bibitem[Cuzzi et al.(1993)]{cuz93}
1490: Cuzzi, J. N., Dobrovolskis, A. R., \& Champney, J. M. 1993, \icarus, 106, 102
1491: \bibitem[Cuzzi \& Hogan(2003)]{cuz03}
1492: Cuzzi, J. N., \& Hogan, R. C. 2003, \icarus, 164, 127
1493: \bibitem[Cuzzi \& Weidenschilling(2006)]{cuz06}
1494: Cuzzi, J. N., \& Weidenschilling, S. J. 2006, In Metoerites of the Early
1495: Solar System II, D. S. Loretta, and H. Y. McSween, Jr., eds.,
1496: (Tucson: Univ of Arizona Press), 353
1497: \bibitem[Dominik et al.(2007)]{dom07}
1498: Dominik, C., Blum, J., Cuzzi, J. N., \& Wurm, G. 2007, In Protostars and
1499: Planets V, B. Reipurth, D. Jewitt, and K. Keil, eds.,
1500: (Tucson: Univ. of Arizona Press), 783
1501: \bibitem[Dullemond \& Dominik(2005)]{dul05}
1502: Dullemond, C. P., \& Dominik, C. 2005, \aap, 434, 971
1503: \bibitem[Durisen et al.(2007)]{dur07}
1504: Durisen, R. H., Boss, A. P., Mayer, L., et al., 2007,
1505: In Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil, eds,
1506: (Tucson: Univ of Arizona Press), 607
1507: \bibitem[Garaud(2007)]{gar07}
1508: Garaud, P, 2007, arXiv:0705.1563, Submitted to ApJ
1509: \bibitem[Hubickyj et al.(2005)]{hub05}
1510: Hubickyj, O., Bodenheimer, P., \& Lissauer, J. J., 2005, \icarus, 179, 415
1511: \bibitem[Johansen et al.(2007)]{joh07}
1512: Johansen, A., Oishi, J. S., Low, M.-M. M., et al. 2007, \nat, 448, 1022
1513: \bibitem[Kessler-Silacci et al.(2006)]{kes06}
1514: Kessler-Silacci, J., Augereau, J.-C., Dullemond, C. P., et al. 2006, \apj,
1515: 639, 275
1516: \bibitem[Kenyon \& Luu(1999)]{ken99}
1517: Kenyon, S. J., \& Luu, J. X., 1999, \apj, 526, 465
1518: \bibitem[Leinhardt \& Richardson(2005)]{lei05}
1519: Leinhardt, Z. M., \& Richardson, D. C, 2005, \apj, 625, 427
1520: \bibitem[Loginov(1979)]{log79}
1521: Loginov, V. I. 1979, Dehydration and Desalinization of Oil: Khimiya, Moscow
1522: \bibitem[Markiewicz et al.(1991)]{mar91}
1523: Markiewicz, W. J., Mizuno, H., \& V\"{o}lk, H. J. 1991, \aap, 242, 286
1524: \bibitem[Marov \& Kolesnichenko(2001)]{mar01}
1525: Marov, M. Ya., \& Kolesnichenko, A. V. 2001, Mechanics of Turbulence of
1526: Multicomponent Gases: Kluwer Academic Publishers
1527: %\noindent
1528: %Mathis, J. S., Rumpl, W., \& Nordsieck, K. H., 1977, ApJ, 217, 425
1529: \bibitem[Nakagawa et al.(1986)]{nak86}
1530: Nakagawa, Y., Sekiya, M., \& Hayashi, C. 1986, \icarus, 67, 375
1531: \bibitem[Natta et al.(2007)]{nat07}
1532: Natta, A., Testi, L., Calvet, N., et al., 2007, In
1533: Protostars and Planets V , B. Reipurth, D. Jewitt, and K. Keil, eds,
1534: (Tucson: Univ of Arizona Press), 767
1535: \bibitem[Ormel \& Cuzzi(2007)]{orm07}
1536: Ormel, C. W., \& Cuzzi, J. N. 2007, \aap, 413, 466
1537: \bibitem[Ormel et al.(2007)]{orm07b}
1538: Ormel, C. W., Spaans, M., \& Tielens, A. G. G. M. 2007, \aap, 461, 215
1539: \bibitem[Ormel et al.(2008)]{orm08}
1540: Ormel, C. W., Cuzzi, J. N., \& Tielens, A. G. G. M. 2008, \aap, submitted
1541: \bibitem[Pollack et al.(1996)]{pol96}
1542: Pollack, J. B., Hubickyj, O., Bodenheimer, P., et al., 1996, \icarus, 124, 62
1543: \bibitem[Press et al.(1992)]{pre92}
1544: Press, W. H., Teukolsky, S. A., Vetterling, W. T., \& Flannery, B. P. 1992,
1545: Numerical Recipes in Fortran: 2nd edition, Cambridge University Press.
1546: \bibitem[Przygodda et al.(2003)]{pry03}
1547: Przygodda, F., van Boekel, R., \`{A}brah\`{a}m, P., et al. 2003, \aap, 412,
1548: L43
1549: \bibitem[Saffman \& Turner(1956)]{saf56}
1550: Saffman, P. G., \& Turner, J. S., 1956, J. Fluids Mech., 1, 16
1551: \bibitem[Safronov(1969)]{saf69}
1552: Safronov, V. S. 1969, Evolution of the Protoplanetary Cloud and the Formation
1553: of the Earth and Planets: NASA TTF-677, 1972
1554: \bibitem[Silk \& Takahashi(1979)]{sil79}
1555: Silk, J., \& Takahashi, T., 1979, \apj, 229, 242
1556: \bibitem[Smoluchowski(1916)]{smo16}
1557: Smoluchowski, M., 1916, Phys. Z, 17, 557
1558: \bibitem[Stepinski \& Valageas(1997)]{ste97}
1559: Stepinski, T. F., \& Valageas, P., 1997, \aap, 319, 1007
1560: \bibitem[Tanaka et al.(2005)]{tan05}
1561: Tanaka, H., Himeno, Y., \& Ida, S., 2005, \apj, 625, 414
1562: \bibitem[Trubnikov(1971)]{tru71}
1563: Trubnikov, B. A. 1971, Soviet Phys. - Doklady, 16, 124
1564: \bibitem[van Boekel et al.(2003)]{van03}
1565: van Boekel, R., Waters, L. B. F. M., Dominik, C., et al. 2003, \aap, 400, L21
1566: \bibitem[V\"{o}lk et al.(1980)]{vol80}
1567: V\"{o}lk, H. J., Morfill, G. E., Roeser, S., \& Jones, F. C. 1980,
1568: \aap, 85, 316
1569: \bibitem[Weidenschilling(1977)]{wei77}
1570: Weidenschilling, S. J., 1977, \mnras, 180, 57
1571: \bibitem[Weidenschilling(1984)]{wei84}
1572: Weidenschilling, S. J., 1984, \icarus, 60, 553
1573: \bibitem[Weidenschilling(1997)]{wei97}
1574: Weidenschilling, S. J., 1997, \icarus, 127, 290
1575: \bibitem[Weidenschilling(2000)]{wei00}
1576: Weidenschilling, S. J., 2000, \ssr, 92, 295
1577: \bibitem[Weidenschilling(2002)]{wei02}
1578: Weidenschilling, S. J., 2002, Met.\& Pla. Sci., 37, A148.
1579: \bibitem[Weidenschilling(2004)]{wei04}
1580: Weidenschilling, S. J., 2004, In Comets II, ed. M. C. Festou, H. U. Keller,
1581: \& H. A. Weaver (Tucson: Univ. of Arizona Press), 97
1582: \bibitem[Weidenschilling \& Cuzzi(1993)]{wei93}
1583: Weidenschilling, S. J., \& Cuzzi, J. N. 1993, In Protostars and Planets
1584: III, ed. E. H. Levy \& J. I. Lunine (Tucson: Univ. of Arizona Press), 1031
1585: \end{thebibliography}
1586:
1587: \clearpage
1588:
1589: \begin{figure}
1590: \plotone{f1.ps}
1591: \caption{Comparison between the scaled integer moments $M^\prime_k =
1592: M_k/M_k(0)$ for $k = 0,1,2$ (curves) obtained
1593: by the method of moments calculation with a
1594: simple Saffman-Turner turbulent coagulation kernel (Eq. 7) and a brute
1595: force integration (symbols) of the coagulation equation (Eq. 1). The integer
1596: moments for the latter are calculated a priori using the distribution function
1597: $f(m,t)$. There is good agreement, especially for the higher resolution
1598: case.}
1599: \end{figure}
1600:
1601: \clearpage
1602:
1603:
1604: \begin{figure}
1605: \plotone{f2.ps}
1606: \caption{Comparison of the scaled integer moments $k=0,1,2$ obtained from
1607: the moments method (curves), and those obtained from the integration of the
1608: coagulation equation (Eq. 1, symbols) for the case of a realistic collision
1609: kernel with
1610: $v_T = 0$. Both the explicit (solid curves), and implicit (short-dashed
1611: curves) approaches match the coagulation calculation fairly well, especially
1612: the higher resolution case.}
1613: \end{figure}
1614:
1615: \clearpage
1616:
1617: \begin{figure}
1618: \plotone{f3.ps}
1619: \caption{Comparison of the scaled integer moments $M_k$ for $k=0,1,2$ obtained
1620: from the moments method, with results obtained from the brute force integration
1621: of the coagulation
1622: equation (Eq. 1) for the case of a realistic collision kernel with
1623: $v_T \neq 0$ (and no systematic velocities). In this case the turbulence
1624: parameter $\alpha = 10^{-4}$. Both the explicit (solid curves) and
1625: implicit (short-dashed curves) approaches have some difficulty matching
1626: the coagulation calculation for both $M_0$ and $M_2$ (note that the solid and
1627: short-dashed curves
1628: for $M_2$ lie on top of each other). However, the semi-implicit approach
1629: (long-dashed curves) provides a better fit. The modified explicit
1630: approach (dotted curves) provides a better fit for $M_2$ while giving a worse
1631: fit for $M_0$ (see {\S} 3.2).}
1632: \end{figure}
1633:
1634: \clearpage
1635:
1636: \begin{figure}
1637: \plotone{f4.ps}
1638: \caption{Comparison of the second moment $M^\prime_2(t)$ obtained from the
1639: moments method with that obtained from integration of the coagulation
1640: equation (Eq. 1)
1641: for the case of a realistic collision kernel with $v_T \neq 0$ (and no
1642: systematic velocities) for three different values of the turbulence
1643: parameter $\alpha$. Both the explicit (solid curves) and implicit
1644: (short-dashed curves) approaches match the lower $\alpha$ values fairly well,
1645: but as
1646: indicated in fig. 3, have difficulty matching the case of $\alpha = 10^{-4}$
1647: (note that the solid and short-dashed curves lie on top of each other).
1648: However, not surprisingly, the semi-implicit approach (long-dashed curves)
1649: and a modified explicit approach (for $M_2$ only, dotted curve) provides a
1650: much better fit (see {\S} 3.2).}
1651: \end{figure}
1652:
1653: \clearpage
1654:
1655: \begin{figure}
1656: \plotone{f5.ps}
1657: \caption{Comparison of particle growth as a function of time using the
1658: explicit approach (Eq. 28, solid curve), and the \citet{gar07} analytical
1659: particle growth expression (long-dashed curve). The upper and
1660: lower sets of curves
1661: differ in the initial size of the largest particle. Both
1662: sets of curves begin to diverge immediately. The somewhat subdued kink in the
1663: upper curves
1664: at $r_L \sim 13.5$ cm is due to a shift from Epstein to Stokes flow.
1665: However, overall, the agreement is not bad ($\sim 20-30$\% in $r_L$).}
1666: \end{figure}
1667:
1668: \clearpage
1669:
1670: \begin{figure}
1671: \plotone{f6.ps}
1672: \caption{Plot of the normalized (to initial value) wavelength-independent
1673: opacities for different choices of the turbulence parameter $\alpha$.
1674: Opacities decrease sharply (in the absence of any mechanism to hinder particle
1675: growth) for higher $\alpha$. Calculations were performed using the
1676: semi-implicit approach with both turbulent and systematic relative
1677: velocities included. }
1678: \end{figure}
1679:
1680:
1681: \end{document}