Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Icarus 180 (2006) 224–242

www.elsevier.com/locate/icarus

Non-gravitational force modeling of Comet 81P/Wild 2


I. A nucleus bulk density estimate
Björn J.R. Davidsson a,∗ , Pedro J. Gutiérrez b
a Department of Astronomy and Space Physics, Uppsala University, Box 515, SE-75120 Uppsala, Sweden
b Instituto de Astrofísica de Andalucía-CSIC, Aptd 3004, 18080 Granada, Spain

Received 23 February 2005; revised 11 July 2005


Available online 19 September 2005

Abstract
The nucleus of Comet 81P/Wild 2 is modeled by assuming various smooth triaxial ellipsoidal or irregular body shapes, having different
rotational periods, spin axis orientations, and thermophysical properties. For these model nuclei, a large number of surface activity patterns (e.g.,
maps of active and inactive areas) are studied, and in each case the resulting water production rate and non-gravitational force vector versus
time are calculated. By requiring that the model nuclei simultaneously reproduce certain properties of the empirical water production curve and
non-gravitational changes of the orbit (focusing on changes of the orbital period and in the longitude of perihelion), constraints are placed on
several properties of the nucleus. The simulations suggest that the mass of Comet 81P/Wild 2 is M  2.3 × 1013 kg, resulting in a rather low bulk
density, ρbulk  600–800 kg m−3 (depending on the assumed nucleus volume), and that the nucleus rotation is prograde rather than retrograde.
The active area fraction is difficult to constrain, but at most 60% of the nucleus is likely to have near-surface ice.
 2005 Elsevier Inc. All rights reserved.

Keywords: Comets; 81P/Wild 2; Dynamics; Bulk density

1. Introduction In a forthcoming paper, the rotational state of the nucleus is ad-


dressed.
On January 2, 2004, the NASA/JPL spacecraft Stardust In the present work, 81P is modeled either as a smooth triax-
made a successful flyby of Comet 81P/Wild 2 (henceforth, 81P) ial ellipsoid, or as a body with large-scale surface irregularities,
at a minimum distance of 236 km. The dust coma environ- rotating around the shortest principal axis. A large number of
ment was investigated and characterized (Kissel et al., 2004; surface activity patterns (i.e., distributions of icy “active ar-
Tuzzolino et al., 2004), and the nucleus was imaged at high eas” and surrounding inactive dust mantles) are selected, based
on their capability of reproducing empirical water production
resolution (Brownlee et al., 2004). Presently, Stardust is head-
rate measurements, assuming certain thermophysical proper-
ing for Earth, carrying with it invaluable samples of cometary
ties of the model nucleus material. For these activity patterns,
dust particles collected in situ, for ground laboratory analysis.
the corresponding time-dependent non-gravitational force vec-
Certainly, the Stardust mission represents an important step for-
tor, acting on the nucleus due to outgassing, is calculated. By
ward for our understanding of cometary nuclei, and thereby, the
requiring that the empirical change of the orbital period per
processes which once formed the Solar System. However, many
apparition (caused by the non-gravitational force), P, is re-
questions remain unclear, and in the present paper we focus on
produced by the model, the nucleus mass and bulk density can
one of these issues—the nucleus mass and bulk density of 81P.
be estimated. The quality of this estimate is evaluated by testing
if the suggested mass and non-gravitational force simultane-
* Corresponding author. Fax: +46 (0)18 4715999. ously are consistent with the empirical change in the longitude
E-mail addresses: bjorn.davidsson@astro.uu.se (B.J.R. Davidsson), of perihelion per apparition ( ). The change in the longitude
pedroj@iaa.es (P.J. Gutiérrez). of the ascending node per apparition Ω is also calculated, al-
0019-1035/$ – see front matter  2005 Elsevier Inc. All rights reserved.
doi:10.1016/j.icarus.2005.07.023
The nucleus bulk density of Comet 81P/Wild 2 225

though the large uncertainty in the empirical Ω prevents us model data (that are very time-consuming to produce; see Sec-
from using it as an observational constraint, except for a lim- tion 3.2). The periods P = 12 h and P = 25 h are used for the
ited number of extreme cases. The effect on the density estimate HECEM.
when varying the spin axis orientation (SAO) and rotational pe- Prior to the Stardust flyby, three attempts were made to es-
riod is also investigated. timate the spin axis orientation of 81P, by studying the jet
In this work, two different thermophysical models are used morphology during the 1997 apparition. Expressing the spin
in order to investigate the model dependence of the results. The axis orientation in the Eulerian angles (obliquity I and ar-
first, called the layer energy absorption model (LEAM), applied gument Φ 1 ), and here arbitrarily assuming that the sub-solar
for the smooth triaxial body only, assumes the ice/dust mixture point was located on the northern hemisphere2 at perihe-
of active areas to be modestly transparent to solar radiation, heat lion, Sekanina (2003) estimated (I, Φ) = (75◦ , 150◦ ), Vasund-
conducting, and highly porous (e.g., sub-surface sublimation, hara and Chakraborty (see Sekanina et al., 2004) obtained
recondensation, and gas diffusion are taking place). The sec- (I, Φ) = (80◦ , 151◦ ), while Farnham and Schleicher (2005) re-
ond, called the heat conduction equation model (HECEM) is port (I, Φ) = (56 ± 7◦ , 167 ± 6◦ ). After the Stardust flyby,
applicable for solid and totally opaque ice/dust mixtures. This Brownlee et al. (2004) (see also Duxbury et al., 2004) and
model does not account for sub-surface sublimation and gas dif- Sekanina et al. (2004) fitted triaxial ellipsoids to the imaged
fusion, but considers solid-state heat conduction (used for all nucleus, and by assuming simple rotation around the shortest
studied nucleus shape models). principal axis they obtained (I, Φ) = (57◦ , 156◦ ) and (I, Φ) =
The observational data used in this study are presented in (55◦ , 150◦ ), respectively.
Section 2. The thermophysical models and related technical Although the rotational axis appears to be fairly well con-
details are treated in Section 3. The results are presented in strained, the sense of rotation around this axis is somewhat
Section 4 and are discussed in Section 5. Finally, a summary more problematic (the SAO could be similar to one of the val-
is made in Section 6. ues quoted above, or having the opposite direction). Sekanina
et al. (2004) suggest that the rotation of 81P is retrograde (i.e.,
2. Observational data I > 90◦ ), by comparing images of dust jets taken by Star-
dust just prior to the closest approach, with the activity burst
2.1. Nucleus size, shape, rotational period, and spin axis time-sequence detected with the Stardust Dust Flux Monitor In-
orientation strument during later stages of the flyby. They therefore prefer
the SAO (I, Φ) = (125◦ , 330◦ ), i.e., their value quoted previ-
Stardust images reveal that 81P approximately is a triax- ously should correspond to the rotational south pole. However,
ial ellipsoid, with semi-axes {a, b, c} = {2.75, 2.00, 1.65} ± Farnham and Schleicher (2005) argue for a prograde rotation
0.05 km (Brownlee et al., 2004; Duxbury et al., 2004). In the (i.e., I < 90◦ ), based on the curvature of jets seen in images
following, these values are used to define the smooth triaxial taken during the 1997 apparition. In the following, we there-
ellipsoid model nucleus applied in this work (and the irregu- fore study both prograde and retrograde orientations, focusing
lar model bodies also roughly have these dimensions). After on (I, Φ) = (56◦ , 153◦ ), and its reverse, (I, Φ) = (124◦ , 333◦ ),
submission of the current paper, a refined image analysis was obtained by averaging the two estimates derived from Stardust
performed by Howington-Kraus et al. (2005), resulting in a new images.
triaxial shape model with {a, b, c} = {2.607, 2.002, 1.350} km,
with errors preliminarily estimated to be “on the 100 m level.” 2.2. Non-gravitational changes of the orbit
We note that all our calculations and presentation of results are
made using the first shape model (although we later estimate Due to conservation of linear momentum during outgassing,
the effect the more recent shape model has on the bulk density a cometary nucleus is subjected to a so-called non-gravitational
estimate). force, which gradually changes the orbital elements with re-
The rotational period is presently unknown. Groundbased spect to the purely gravitational orbit. In the Marsden et
observations of the nucleus have not revealed any obvious mag-
nitude variations (Farnham and Schleicher, 2005), which is un- 1 The obliquity I ∈ [0◦ , 180◦ ] is the angle between the orbital and equato-
derstandable with hindsight, since a nucleus with the previously rial planes of the comet, or equivalently, the angle between the angular mo-
mentioned dimensions, rotating around its shortest principal mentum vector of the orbit and the spin vector of the comet. The argument
axis, shows rather small cross section variations during rota- Φ ∈ [0◦ , 360◦ [ is measured from the vernal equinox of the comet, counter-
clockwise, to its sub-solar meridian at perihelion (Sekanina, 1981), or equiv-
tion. Sekanina et al. (2004) suggest that the rotational period
alently, is the angle measured clock-wise from the negative velocity vector of
is P = 12–24 h, based on an empirical estimate of the product the comet at perihelion, to the projection of the spin vector onto the orbital
P Vd (where Vd is the terminal velocity of dust jet particles), plane.
and assuming plausible values for Vd . Therefore, we here adopt 2 We here apply the following nomenclature: The angular momentum vector

two different rotational periods, P = 12.3 h and P = 25 h for due to nucleus self-rotation defines the spin axis orientation (SAO), or equiv-
the LEAM, in order to investigate the sensitivity of our result alently the “north pole” or “positive pole,” while the opposite direction is the
“south pole” or “negative pole.” An equatorial plane (perpendicular to the an-
on the applied rotational period. These values were used in gular momentum vector and passing through the center of mass) divides the
previous investigations of other comets (Davidsson and Gutiér- body in two—the “northern hemisphere” (containing the north pole) and the
rez, 2004, 2005), which enables us to recycle thermophysical “southern hemisphere” (containing the south pole).
226 B.J.R. Davidsson, P.J. Gutiérrez / Icarus 180 (2006) 224–242

al. (1973) formalism, this force is parameterized by F̄ = lently, P = 11 ± 6 min) is used here as an empirical con-
{Fr , Ft , Fn } = M{A1 , A2 , A3 }g(rh ), where Fr , Ft , and Fn are straint.
the radial, transverse, and normal force components, respec- Fig. 1 (middle panel) shows consistent evidence for a grad-
tively, M is the mass, A1 , A2 , and A3 are constants, and where ual decrease of the longitude of perihelion over time, and in
the force magnitude dependence on heliocentric distance rh is the following,  = −2.0 ± 0.5 is adopted. Unfortunately,
given by the estimates of Ω span a much wider range (as seen in
−2.15  5.093 −4.6142
Fig. 1, lower panel), taking both positive and negative values,
 
rh rh i.e., it is not even clear whether the longitude of the ascend-
g(rh ) = 0.111262 1+ . ing node increases or decreases with time. Averaging of the
2.808 2.808
(1) nine estimates shown in the figure yields Ω ≈ 0.1 , how-
ever, due to the possible change in the non-gravitational pa-
This activity function is symmetric around perihelion, which
rameters mentioned above, it may be motivated to focus on
limits its applicability for comets with a strong perihelion
estimates for which the observational arcs start not earlier than
force asymmetry. Therefore, Yeomans and Chodas (1989) in-
1988. This leaves us with five estimates with an average of
troduced a modified g-function which peaks τ days before or
Ω ≈ −1.0 . It should also be noted that the only estimates
after perihelion [rh (t) is replaced by rh (t − τ ) in Eq. (1)].
including astrometry from the Stardust optical navigation cam-
The non-gravitational constants {A1 , A2 , A3 , τ } are estimated
era yield Ω ≈ −4.2 and −0.6 , respectively. We therefore
by minimizing the residual between calculated positions and
suspect that Ω in fact is negative (at the time around the
observed astrometry, preferentially covering at least three ap-
1997 apparition). Although this is not used as an empirical con-
paritions.
straint in the present work, we feel that model nuclei yielding
With the empirical non-gravitational parameters at hand, the
Ω < 0 deserve some extra attention and discussion. In any
corresponding changes in the orbital period, longitude of peri-
case, model nuclei yielding a Ω which is substantially out-
helion, and longitude of the ascending node (per apparition) can
side the [−5 , 5 ] region are here considered rather unlikely
be calculated by using (Sekanina, 1993)
representations of 81P.
√  P P 
6π 1 − e2 H e Ft 2.3. The water production rate
P = Fr sin ν dt + dt , (2)
Mn2 q(1 + e) rh
0 0
Apart from three water production rate estimates from 1984
√  P  
q(1 + e)H 1 q(1 + e) (not considered here), all available data are from the 1997 ap-
 = Fr 1 − dt parition. The data consists of three fairly large sets, derived
Mke e rh
0 from spectroscopy of the 630 nm OI1 D line (Fink et al., 1999),
P    hydrogen Lyman-α observations at 121.6 nm made by the
rh SWAN instrument on board the SOHO spacecraft (Mäkinen
+ Ft 1 + sin ν dt , (3)
q(1 + e) et al., 2001), and narrowband photometry of OH at 309 nm
0 (Farnham and Schleicher, 2005). In addition, there is one data
P point obtained by Crovisier et al. (2002), derived from radio
H
Ω = √ Fn rh sin(ω + ν) dt. (4) observations of the 18 cm OH line. Note that the data from
Mk q(1 + e) sin i Mäkinen et al. (2001) recently has been re-calibrated (Mäki-
0
nen, 2004, personal communication), and that the updated rates
In these equations, e is the eccentricity, n [day−1 ] is the are used here.
mean motion, ν is the true anomaly, q [AU] is the perihelion Fig. 2 shows that large systematic differences exist between
distance, P = (q/(1 − e))3/2 [yr] is the orbital period (recal- the sets, e.g., rates based on narrowband photometry and SOHO
culated to days), i is the inclination, ω is the argument of per- data differ a factor 3. When considering one set at a time, this
ihelion, and k is the Gaussian gravitational constant. By using consistently reveals a strong perihelion asymmetry—the water
the conversion factor H = 0.049900175 AU s2 m−1 day−2 , SI production rate peaks several weeks before perihelion, which
units can be used for M and F̄ , while P is measured in days, is a tell-tale sign of a substantially non-spherical nucleus shape
whereas  and Ω are measured in radians (recalculated to and/or a strong concentration of ice-rich material to a few iso-
arc seconds throughout this paper). lated regions on the nucleus. Farnham and Schleicher (2005)
Fig. 1 shows estimates of P,  , and Ω based on estimate that the production rate peaks around T = −80 days
available sets of {A1 , A2 } marked as crosses, and on sets of (T denotes the number of days after perihelion), while Fink
{A1 , A2 , A3 , τ } marked as squares. As pointed out by Sekanina et al. (1999) obtain their highest rates for measurements made
(2003), a change in the non-gravitational parameters appears at T ≈ −65 and −37 days. In addition, the visual lightcurves
to have taken place around 1990, resulting in a reduction of in 1990 and 1997 peaked at T ≈ −63 and −21 days, respec-
P (upper panel). Since the water production rate data (Sec- tively (Sekanina, 2003), although the lightcurve may be more
tion 2.3) used in this study are from 1997, it is relevant to con- sensitive to the amount of dust currently in the coma, the dust
sider P estimates for which the observational arc is centered production rate, dust size distribution, and observational geom-
on that year, therefore P = 0.0076 ± 0.0045 days (equiva- etry (i.e., the phase angle), than to the water production rate.
The nucleus bulk density of Comet 81P/Wild 2 227

Fig. 1. Empirical estimates of P (change in the orbital period per apparition; upper panel),  (change in the longitude of perihelion per apparition; middle
panel), and Ω (change in the longitude of the ascending node per apparition; lower panel), calculated from sets of {A1 , A2 } (crosses) or sets of {A1 , A2 , A3 , τ }
(squares). Horizontal bars indicate the observational arc, where symbols mark the midpoint. Sources: MPC 40671, Nakano (OAA computing section circular
NK 716), Kinoshita (URL http://www9.ocn.ne.jp/comet/pcmtn/0081p.htm), Rocher (Cometary Notes of Bureau des Longitudes, No. 0008), JPL (#28, #J974/1,
#K033/7, #K033/47, #C197/1), S. Chesley (2003, 2004, personal communication), and S. Szutowicz (2005, personal communication).

The fact that the peaks in the measured water production rate, 3. Method
and the visual lightcurve, both take place pre-perihelion may
therefore be a coincidence. The non-gravitational force (possi- 3.1. Thermophysical models
bly strongly correlated to the water production rate) may have
reached its maximum value around T = −66 days, based on As mentioned previously, two thermophysical models are
an average τ , obtained by considering nine estimates. Taken applied in the current work. These are described briefly in the
together, this suggests a peak in the production rate at T  following.
−40 days, which is used in the following as an empirical con-
3.1.1. The layer energy absorption model, LEAM
straint.
The LEAM considers the nucleus material as a porous, par-
The quality of a bulk density estimate depends strongly on
tially transparent and heat conducting mixture of crystalline
the empirical water production rates used in the analysis, since water ice and sub-µm dust grains. Thermal coupling between
these are necessary in order to evaluate the active area fraction ice and dust is assumed, i.e., the two substances have the same
of the nucleus and, to some extent, constrain the latitudinal dis- temperature. Due to the structural, optical, and thermal proper-
tribution of active areas. Due to the substantial uncertainties in ties of such a medium, a number of features must be taken into
the observed water production rates for 81P, an accurate bulk account in the model. Solar energy is not absorbed merely at the
density estimate is difficult to make. However, as it turns out, surface, but is deposited gradually inside a surface layer (hence
it is still possible to obtain a non-trivial upper limit on the bulk the name of the model). The existence of pore spaces and chan-
density for the comet, by defining a production rate envelope nels inside the material leads to sub-surface ice sublimation,
which encapsulates the available measurements with a rather vapor recondensation and gas diffusion, processes which serve
substantial margin. The envelope used in the present investiga- as energy sources and sinks, as well as transport mechanisms
tion is shown as dashed curves in Fig. 2. for heat and mass.
228 B.J.R. Davidsson, P.J. Gutiérrez / Icarus 180 (2006) 224–242

Fig. 2. The measured water production rate of Comet 81P/Wild 2 for the 1997 apparition (dashed curves show an envelope which approximately encapsulates the
observed data).

The LEAM has been described in a number of papers, there- t and depth x, solid-state heat conduction, convection (i.e.,
fore only a brief summary is given here. The first version of energy carried by diffusing gas molecules within the porous
the model was described in detail by Davidsson and Skorov medium), absorption of energy from the sub-surface solar ra-
(2002b), and preparatory work on light absorption in porous diation field, and net changes in energy content of the solid
media was presented by Davidsson and Skorov (2002a). The medium due to sublimation or condensation. The terms in the
model was extended to deal with a quasi-parallel simulation of mass conservation equation, Eq. (6), denote (again, from left to
the nucleus and the overlaying coma by Davidsson and Skorov right), the total change in gas number density ρg (x, T ) versus
(2004). The full model was also summarized by Davidsson and time and depth, change in number density due to gas diffusion,
Gutiérrez (2004, 2005). and change in number density due to net sublimation or con-
Two spatially one-dimensional and coupled differential densation. In these equations, ψ is the porosity, ρi and ρd are
equations describing energy and mass conservation govern the the densities of solid ice and dust, respectively, fi and fd are
thermophysical behavior of the model nucleus, the volumetric fractions of ice and dust in compacted mater-
   ∂T (x, t) ial, respectively, while the heat capacities are ci (T ) (ice), cd (T )
(1 − ψ) ρi fi ci (T ) + ρd fd cd (T ) (dust), and cg (vapor). The conductivity is denoted by κ(T ) [see
∂t
  Eq. (12)], while φg (ρg , T ) is the mass flux rate (i.e., the flux
=

κ(T )
∂T (x, t)
− cg φg (ρg , T )
∂T (x, t) of mass [kg m−2 s−1 ] inside the medium due to gas diffusion),
∂x ∂x ∂x and qm (ρg , T ) is the volume mass production rate [kg m−3 s−1 ]
  ∂F (x) (i.e., the net amount of gas molecules being created or removed
− I A, rh , z(t) − qm (ρg , T )L(T ), (5) by sublimation and condensation, respectively). The solar flux
∂x
as function of depth below the surface is given by F(x) (here,
∂ρg (x, t) ∂φg (ρg , T ) Beer’s law is used, normalized to unity at the surface), and
ψ =− + qm (ρg , T ). (6)
∂t ∂x I(A, rh , z(t)) is the absorbed solar flux as function of albedo A,
The terms in the energy conservation equation, Eq. (5) (from heliocentric distance rh and zenith angle z (which varies with
left to right), account for changes in temperature T versus time time during nucleus rotation).
The nucleus bulk density of Comet 81P/Wild 2 229

The upper and lower boundary conditions for Eq. (5) are face normal n̂m , due to outgassing, is given by
given by
  F̄m = −p(Ts , T )am n̂m , (10)
εσ Ts4 + (1 − ψ)fi 1 − B(Ts , T ) as (Ts )ρsat (Ts )
 where p(Ts , T ) is the pressure acting on the surface (recoil
k B Ts ∂T dIF and impact contributions from sublimating, recondensing, and
× L(Ts ) − κ(Ts ) = δx, (7a)
2πmH2 O ∂x x=0 dx x=0 scattered molecules).

∂T Explicit expressions for heat capacities, conductivity, la-
= 0. (7b)
∂x x=xb
tent heat, and saturation density are given by Davidsson and
Skorov (2002b). The quantities B(Ts , T ), D(Ts , T ), and
Equation (7a) balances thermal reradiation into space, net sur- p(Ts , T ) have been obtained by performing Direct Simu-
face sublimation, heat conduction (generally from the warmer lation Monte Carlo modeling of the cometary Knudsen layer
interior to the surface, due to the solid-state greenhouse effect), vapor (Davidsson and Skorov, 2004), and that paper also con-
with the small amount of solar radiation absorbed in an infi- tains explicit expressions for the mass flux rate, volume mass
nitely thin surface slab. Equation (7b) states that the energy production rate, sublimation coefficient, and condensation co-
flux across an internal surface located at depth xb is negligible. efficient. For a detailed discussion about F , see Davidsson and
This is a reasonable assumption as long as the heliocentric dis- Skorov (2002b) [the current treatment of F is furthermore iden-
tance is not larger than ∼3 AU (beyond this distance, heat losses tical to that by Davidsson and Gutiérrez (2004)]. The values of
through conduction to the deep interior becomes more impor- all parameters used in the LEAM have been summarized in Ta-
tant than losses due to near-surface sublimation). Typically, xb ble 1.
is on the order of a few meters in the current work, since diurnal
temperature variations are non-existent at such depths. In these 3.1.2. The heat conduction equation model, HECEM
equations, ε is the emissivity, σ is the Stefan–Boltzmann con- In this model, the (thermally coupled) ice/dust mixture is
stant, B(Ts , T ) is the fraction of sublimated molecules which assumed to be solid (but potentially containing small uncon-
return to the surface from the coma and condense (a function nected void spaces), totally opaque, and heat conducting. The
of the surface temperature and the temperature difference T HECEM is a modification of the EBEM (energy balance equa-
over a thin surface slab), as (T ) is the sublimation coefficient (an tion model) used by Davidsson and Gutiérrez (2005), since that
empirical correction to the classical Hertz–Knudsen formula), model did not take conductivity into account. Since no net sub-
ρsat (T ) is the saturation density, kB is the Boltzmann constant, surface sublimation or gas diffusion can take place in a solid
mH2 O is the water molecule mass, and L(T ) is the latent heat medium, and since all light absorption occurs at the very sur-
of ice. face, the single governing equation is given by
The upper and lower boundary conditions for Eq. (6) are  
given by   ∂T (x, t) ∂ ∂T (x, t)
ρi fi ci (T ) + ρd fd cd (T ) = κ(T ) .
∂t ∂x ∂x
ρg |x=0 = D(Ts , T ), (8a) (11)
as (Tb )ρsat (Tb ) The conductivity is given by the same expression as for the
ρg |x=xb = , (8b)
ac (Tb ) LEAM, i.e.,
where D(Ts , T ) is the gas density at the nucleus/coma bound-  
567fi
ary, and ac (T ) is the condensation coefficient (the fraction of κ(T ) = h + fd κd , (12)
impinging molecules that actually condense on an icy surface, T
instead of being scattered). where h is the Hertz factor, which is unity for a solid ice/dust
Once the time-dependent temperature and gas density struc- mixture, but may be substantially smaller if unconnected “bub-
tures of the medium have been obtained by solving Eqs. (5)– bles” are present. The heat capacity of ice is given by
(10), the total backflux-corrected water production rate (from
the surface and the near-surface interior) is given by ci = 7.49T + 90 (13)


  k B Ts (note that SI units are used).
Z = 1 − B(Ts , T ) (1 − ψ)fi as (Ts )ρsat (Ts ) The upper boundary condition for Eq. (11) is given by
2πmH2 O
 
xb I A, rh , z(t) = εσ Ts4 + fi (1 − Bc )as (Ts )ρsat (Ts )
3(1 − ψ)fi  
+ exp 2.0 × 10−3 (x/rg + 1)Ts
rg k B Ts ∂T
× L − κ(Ts ) , (14)
∂x x=0
0
 2πmH2 O
 kB T
− 0.92x/rg − 1 ρsat (T )as (T ) dx , (9) where a constant molecular backflux Bc is assumed [see
2πmH2 O
Davidsson and Skorov (2004) for a discussion about the sub-
where rg is the effective grain radius of the medium. The force limation coefficient as (T )]. The lower boundary condition is
acting vertically on a surface element m with area am and sur- identical to Eq. (7b), with xb = 2.5 m.
230 B.J.R. Davidsson, P.J. Gutiérrez / Icarus 180 (2006) 224–242

Table 1
Summary of applied model parameters
Description Symbol Value Unit
Perihelion distance q 1.591 AU
Eccentricity e 0.539 –
Argument of perihelion ω 41.48 ◦
Longitude of the ascending node Ω 136.14 ◦
Inclination i 3.24 ◦
Orbital period P 6.41 yr
Gaussian gravitational constant k 0.017202099 AU3/2 day−1
Conversion factor H 0.049900175 AU s2 m−1 day−2
Change of orbital period P 0.0076 ± 0.0045 Days
Change in longitude of perihelion  −2.0 ± 0.5 arc sec
Rotational period P 12 (alt. 12.3) or 25 h
Nucleus dimensions 2.75 × 2.00 × 1.65 km
Stefan–Boltzmann constant σ 5.6705 × 10−5 W m−2 K−4
Boltzmann constant kB 1.38066 × 10−23 J K−1
Water molecule mass mH2 O 2.98897 × 10−26 kg
Volume fraction of ice (compact material) fi 0.731 –
Volume fraction of dust (compact material) fd 0.269 –
Emissivity ε 0.97 –
Plane albedo A 0.032 –
LEAM porosity ψ 0.7 –
LEAM compact ice density ρi 933 kg m−3
LEAM compact dust density ρd 2078 kg m−3
LEAM e-folding scale ζ 1.95 × 10−2 m
LEAM dust conductivity κd 3.1 W m−1 K−1
LEAM Hertz factor h 0.1 –
LEAM tortuosity ξ 1 –
LEAM channel length x 10−2 m
LEAM grain radius rg 10−2 m
LEAM pore radius rp 10−2 m
HECEM compact ice density ρi 917 kg m−3
HECEM compact dust density ρd 3500 kg m−3
HECEM latent heat of ice L 48600 J mol−1
HECEM dust heat capacity cd 1200 J kg−1 K−1
HECEM dust conductivity κd 4.2 W m−1 K−1
HECEM Hertz factors h 0, 0.01, or 0.5 –
HECEM backflux Bc 12 %
HECEM momentum transfer coefficient η 0.77 –
Note. Various parameters used in the modeling (common to both LEAM and HECEM if not stated otherwise). Orbital elements are given for Equinox J2000.0. Note
that the values of fi and fd are derived from the mass ratios of silicates, organics, ice, and carbon suggested by Greenberg and Hage (1990) as typical for comet
material.

Once the surface temperature has been calculated, the water are similar to the ones used by Gutiérrez et al. (2001), i.e., they
production rate is evaluated as have Gaussian random shapes, starting out with the same shape
 parameter values as used in that work. However, the princi-
Z = fi (1 − Bc )as (Ts )ρsat (Ts )
k B Ts
, (15) pal axes have subsequently been adjusted, so that a/b ≈ 1.38
2πmH2 O and a/c ≈ 1.67 (as for 81P), and then these bodies have been
slightly re-scaled to yield the same volume as for the smooth tri-
and the force acting on a surface element due to outgassing is
axial ellipsoid (so that the derived bulk densities become com-
approximated by
parable). We note that these irregular bodies are not intended as

8kB Ts accurate representations of the real shape of 81P—we merely
F̄m = −ηZV am n̂m = −ηZ am n̂m , (16) want to investigate the sensitivity of our results with respect to
πmH2 O
shape. Table 2 shows some characteristics of the irregular bod-
where η is the momentum transfer efficiency (see Table 1 for ies, compared to the smooth triaxial model nucleus. Drawings
further information). of the bodies can also be seen in Fig. 3.
The surface of an irregular body is divided into 1520 locally
3.2. Practical and technical details flat triangular facets, approximately extending 9◦ in longitude
and latitude. The surface of the smooth triaxial model body is
In this work, one smooth triaxial ellipsoid and three differ- divided into 3538 locally flat triangular facets, each extending
ent irregular model bodies are considered. The irregular bodies 5.9◦ in longitude and 6.0◦ in latitude.
The nucleus bulk density of Comet 81P/Wild 2 231

Fig. 3. Drawings of the smooth triaxial ellipsoid, and the three irregular nucleus shape models used in this study.

Table 2
Properties of irregular bodies
Body Highland (%) Mean height (m) Max height (m) Lowland (%) Mean depth (m) Max depth (m)
Irr#1 44 202 791 56 178 619
Irr#2 42 124 634 58 248 593
Irr#3 48 110 411 52 143 440
Note. Properties of three irregular bodies used in the study (see also Fig. 3). A smooth triaxial ellipsoid with semi-axes {a, b, c} = {2.75, 2.00, 1.65} km is used as a
zero-elevation surface. Column 2 shows the percentage of the irregular surface above zero-elevation, while columns 3 and 4 show the mean and maximum heights
of the “highlands.” Columns 5–7 show corresponding data for the “lowlands.”

In the present work, it is postulated that the comet nu- and for each area generating a random position vector extending
cleus consists of two fundamentally different types of surface from the body center to the surface. Next, cones with arbitrary
materials—potentially active areas, which have exposed sur- angular widths (60◦ ), centered on the position vectors are
face ice mixed with dust (producing vapor according to the formed, and all facets within a cone are considered potentially
thermophysical models both day and night, except when be- active.
ing in constant darkness during polar nights), and inactive areas Calculations are performed for 131 points in the orbit, us-
which lack surface ice, i.e., possible sub-surface icy material ing a 10-day time step when 2 < rh  3 AU and a 3-day
is covered by a thick dust mantle (assumed not to emit vapor time step when rh  2 AU (the total simulated time period
under any illumination condition). Furthermore, these units are is 600 days, centered on perihelion). For each orbital posi-
assumed stable over time, e.g., dust mantling or mantle blow- tion, 12 discrete rotational phases are considered in order to
off has not been considered. calculate daily averages of the water production rate and the
Each facet is considered either potentially active or inactive, non-gravitational force vector (however, the marching time-step
and a certain surface distribution of such facet types constitutes during nucleus rotation, when actually solving the thermophys-
a “surface activity pattern” (or trial map). A trial map is con- ical equations, is on the order of 1 s for the LEAM and 100 s
structed by selecting an arbitrary number (1–10) of active areas, for the HECEM).
232 B.J.R. Davidsson, P.J. Gutiérrez / Icarus 180 (2006) 224–242

To solve the thermophysical equations presented in Sec- Section 2.1 therefore means that we here apply previously cal-
tion 3.1 for a single surface facet on a model nucleus located in culated thermophysical look-up tables obtained for spin periods
a particular point in the orbit (having a certain SAO), in order P = 12.3 h and P = 25 h, for the material parameters shown in
to obtain the temperature structure and water production rate Table 1.
versus time during nucleus rotation, is a rather time-consuming Finally, we note that the HECEM, applied for the irregular
process. Doing this for thousands of facets, at a large number bodies, fully takes local shadowing due to surface topography
of points in the orbit, is clearly infeasible if a head-on approach into account. However, self-heating (arising when parts of the
is applied, especially given the number of surface activity pat- surface absorbs thermal radiation emitted by other regions) has
terns and spin axis orientations considered here. Therefore, a not been considered, since Gutiérrez et al. (2001) showed that
technique for optimizing calculations of this kind, while still it generally has a very small effect on the energy budget of the
preserving accuracy, has been developed. The interested reader nucleus.
may consult Davidsson and Gutiérrez (2004, 2005), where the
technique is described in detail. In short, the method consists of 4. Results
fixing a certain rotational period and set of material parameters.
A large number of thermophysical models are then calculated, Before presenting the results for spotted model nuclei, it is
to cover a wide range of time-dependent illumination condi- important to mention some characteristics of the water pro-
tions. The collected data form a look-up table, which can be duction curve of a fully active model nucleus (i.e., having no
used to accurately map surface temperatures, water production inactive regions), since it facilitates the understanding of the
rates, and non-gravitational force vectors onto nuclei of arbi- subsequent discussion. Fig. 4 shows the calculated daily mean
trary size, shape, spin axis orientation, and orbital position, tak- water production rate Qcalc (solid curve), as well as the mini-
ing local illumination conditions and time of day into account mum and maximum daily production rates during nucleus rota-
(including night-time activity). The “recycling” mentioned in tion (dashed curves), for a smooth triaxial model nucleus having

Fig. 4. The calculated water production rate of a fully active model nucleus (i.e., no inactive areas) with (I, Φ) = (56◦ , 153◦ ) and P = 12.3 h, using the LEAM
for a smooth triaxial ellipsoid. The solid curve shows the daily mean production rate, while the dashed curves show the daily minimum and maximum productions,
depending on the rotational phase. The dots show the nominal water production rate measurements.
The nucleus bulk density of Comet 81P/Wild 2 233

(I, Φ) = (56◦ , 153◦ ) and P = 12.3 h, using the LEAM. The 81P cannot withstand rotational disruption, if the period of 81P
nominal water production rate measurements from Fig. 2 are indeed is as short as 12 h, and if the body is nearly strengthless].
also shown, as dots. Given the mass, Eqs. (3) and (4) are used to estimate  and
Pre-perihelion, the northern hemisphere is facing the Sun Ω in each case (to be compared with empirical estimates).
(with this assumption regarding the SAO), and the water pro- A total of 103 surface activity patterns are selected for final
duction rate increases comparably steeply with time (Qcalc ∝ study (per combination of SAO, P , nucleus shape and thermo-
rh−5.02 ). This is not only due to the gradual solar flux intensi- physical model).
fication, but because the sub-solar point is climbing northward
(which increases the mean illuminated cross section of the non- 4.1. Smooth triaxial model nuclei studied with the LEAM
spherical nucleus). The summer solstice of the northern hemi-
sphere occurs about 120 days before perihelion, i.e., this is the Fig. 5 shows  versus the nucleus bulk density ρbulk , for
time when the largest fraction of the nucleus is illuminated (the the case (I, Φ) = (56◦ , 153◦ ) and P = 12.3 h, obtained for
sub-solar point reaches the latitude 56◦ N). The water produc- 103 surface activity patterns, using the LEAM for the smooth
tion rate continues to rise beyond this point since the heliocen- triaxial body [note that the nucleus dimensions obtained by
tric distance is decreasing, but as the sub-solar point steadily Brownlee et al. (2004) and Duxbury et al. (2004) are used con-
creeps back towards the equator, the increase in the solar flux sistently throughout Section 4]. Filled circles are used to denote
is insufficient to compensate for the diminishing mean illumi- Ω < 0 (perhaps the most realistic cases), while open circles
nated cross section. This leads to a water production rate peak denote Ω  0 .
around 45 days before perihelion (being 20% higher than at A broad variety of theoretical water production rate curves
the smallest heliocentric distance of the orbit). The subsequent can be fitted within the envelope, hence the properties of the
drop in Qcalc is rather steep at first, but around 45 days post- surface activity maps and the corresponding non-gravitational
perihelion, the sub-solar point is crossing the equator, entering force vectors vary substantially, and the resulting bulk densities
the southern hemisphere (note that equatorial illumination leads cover a very wide range (Fig. 5 has been truncated at ρbulk =
to a substantial daily variation in the water production, as seen 2000 kg m−3 , but densities twice as high can be found). The
in Fig. 4). The gradual increase of the mean illuminated cross highly scattered empirical water production rates and the P
section following this event, as the southern hemisphere re- criterion are therefore not sufficient to constrain the density,
ceives more light, works very efficiently against the decrease but the additional constraint  ∈ [−2.5 , −1.5 ] leads to a
in solar flux, hence the post-perihelion branch is less steep than non-trivial upper limit on the density, ρbulk  540 kg m−3 . Si-
the pre-perihelion branch (Qcalc ∝ rh−2.43 ), and the water pro- multaneous reproduction of P and  therefore rules out the
duction rate remains high for an extensive period. In fact, a fully high-density tail of the distribution seen in Fig. 5, thereby show-
active model nucleus produces almost four times more water ing that both parameters are equally important in the analysis.
3 AU post-perihelion, than at the same distance pre-perihelion We note that the nominal P has been used to calculate the nu-
(solely due to differences in the nucleus orientation with respect cleus mass and ρbulk for each model nucleus shown in Fig. 5,
to the Sun). but that variation of P within its error bars does not mod-
The introduction of surface activity maps is expected to ify the shape of the distribution in that figure and hence not
modify the previously described water production rate curve the estimated upper limit on ρbulk . For the considered spin axis
substantially. Since different hemispheres are illuminated pre- orientation, Ω takes both positive and negative values (also
and post-perihelion, the production rate asymmetry around per- for the most interesting  region), depending on the sur-
ihelion is expected to be highly dependent on the relative degree face activity pattern (and ρbulk  540 kg m−3 holds also for the
of ice coverage of the hemispheres. For example, if ice is much Ω < 0 cases).
more abundant in the north than in the south, the asymmetry It is important to study the reasons why some model nuclei
seen in Fig. 4 could increase significantly. Alternatively, if the manage to reproduce the empirical  , while others fail. The
southern hemisphere has more ice coverage, this could reduce former model nuclei may share some properties (e.g., in terms
the pre-perihelion water production dominance, or even shift of the surface activity maps and the resulting water production
the peak production to the post-perihelion branch. rates), that the latter lack. By identifying these properties, addi-
Using the map generation method described in Section 3.2, tional information about 81P may be extracted.
a large number of surface activity maps are constructed, and For this purpose, the model nuclei have been divided into
those resulting in theoretical water production rate curves con- four groups based on the  value: (1)  > −0.5 ,
fined within the envelope shown in Fig. 2 are selected for further (2) −1.5 <   −0.5 , (3) −2.5    −1.5 (i.e., the
study. This is done for different combinations of rotational pe- empirical range), and (4)  < −2.5 . Relevant diagnostic
riods, spin axis orientations, thermophysical models and model tools are the active area fraction of the northern and south-
nucleus shape types. The corresponding non-gravitational force ern hemispheres (AN and AS , respectively, both 100%), the
vectors are calculated, and Eq. (2) is used in each case to es- ratio AS /AN , the day when the water production rate peaks,
timate the mass and density (by applying the empirical P). Tpeak , and the water production curve asymmetry Qasym . The
A second selection is made, so that only test nuclei with ρbulk > latter parameter is here defined as the ratio between the mean
100 kg m−3 are considered [equations derived by Davidsson production during the time intervals −100  T  0 days and
(2001) show that a less dense body with the size and shape of 100  T  200 days. This definition has been chosen since
234 B.J.R. Davidsson, P.J. Gutiérrez / Icarus 180 (2006) 224–242

Fig. 5. Calculated change in the longitude of perihelion per apparition ( ) versus nucleus bulk density (ρbulk ) for smooth triaxial model nuclei with P = 12.3 h
and (I, Φ) = (56◦ , 153◦ ), obtained by using the LEAM. Note that the nucleus volume consistent with the shape model by Brownlee et al. (2004) and Duxbury et al.
(2004) has been applied here. Filled circles indicate a decreasing longitude of the ascending node over time (Ω < 0 ), while open circles are used for Ω  0 .
The solid lines mark the adopted empirical  range.

different hemispheres are illuminated during these time in- AS = 1.1AN , i.e., Group 1 has AS /AN  1.1. It means that
tervals (Qasym therefore partially reflects differences in ice the active area fraction of the southern hemisphere is always
coverage between the hemispheres). For a fully active nu- higher, or much higher, than that of the northern hemisphere,
cleus, Qasym = 3.8, and Tpeak = −45 days. If AS /AN  1, for Group 1 nuclei. Since the sub-solar point enters the south-
then Qasym  3.8 is expected, and if AS /AN  1 it is likely ern hemisphere around T = 45 days, high production rates are
that Qasym  3.8. However, this should be considered trends or therefore expected post-perihelion. Indeed, Fig. 7 shows that
tendencies rather than rules, since Qasym is also affected by ad- the production rate asymmetry is low, Qasym  2, and in some
ditional factors (such as the individual values of AN and AS , as cases even below unity—the high AS values creates a peak after
well as the exact placement of the active areas). perihelion. As seen in Fig. 7, the water production peak takes
Fig. 6 shows AS versus AN for the 103 model nuclei, and place more than 100 days after perihelion in extreme cases.
Fig. 7 shows Qasym versus Tpeak . In both figures, the four A large fraction of the surface activity patterns for Group 1
groups are distinguished by having different colors. In the fol- nuclei can therefore be considered unlikely, not only for their
lowing, the properties of the four groups are described. inability to reproduce  , but also because the water produc-
tion rates are inconsistent with the observed rates, which seem
Group 1 (red symbols) to peak around Tpeak  −40 days, as mentioned previously.
Model nuclei classified as Group 1 have  > −0.5 , and
are responsible for the highest mass or density estimates in the Group 2 (yellow symbols)
sample (see Fig. 5), although some ρbulk < 540 kg m−3 cases This group is defined by −1.5 <   −0.5 , and con-
also can be found. As all test nuclei in the sample, Group 1 tains model nuclei with bulk densities as high as ρbulk =
nuclei have 15%  AN  60%, but as seen in Fig. 6, red sym- 1500 kg m−3 . In Fig. 6, these cases are concentrated to the re-
bols are approximately separated from the others by a line gion 0.9  AS /AN  1.1, i.e., they represent model nuclei with
The nucleus bulk density of Comet 81P/Wild 2 235

Fig. 6. Active area fraction of the southern hemisphere plotted versus that of the northern hemisphere. Here, 103 surface activity maps are considered, covering a
smooth triaxial ellipsoidal nucleus with (I, Φ) = (56◦ , 153◦ ) and P = 12.3 h. The LEAM was used in all cases.

small differences in ice coverage between the hemispheres. crosses) have more asymmetric water production rate curves,
Fig. 7 shows that Group 2 is clearly different from Group 1 in which peak closer to perihelion. Model nuclei with Ω < 0
the sense that the production rate asymmetry generally is some- and Tpeak  −40 days, have 3  Qasym  4, and a tightly con-
what higher (2  Qasym  4), and that the production rate peak strained AS /AN ratio of about 0.7–0.8, further illustrating that
is restricted to the pre-perihelion branch. the northern active area fraction may be higher than the south-
ern one. But even for this sub-set of model nuclei, the sepa-
Group 3 (black symbols) rate values of AN and AS are rather poorly constrained (e.g.,
Naturally, this is the most interesting group, since these 20%  AN  60%).
model nuclei fulfill the empirical constraint −2.5    Can anything be said about the probable (latitudinal) loca-
−1.5 . As stated previously, only cases with ρbulk  540 kg m−3 tion of active areas on the hemispheres? First, it must be noted
are found. Fig. 6 shows that the majority of the test nuclei are that the concept “location” only is relevant if AN and AS are
located below the line AS = 0.9AN (in particular if Ω < 0 ), fairly small, therefore only cases with AN , AS  30% are con-
i.e., the active area fraction of the southern hemisphere is gen- sidered in the following. Here, the mean latitude is defined as
erally lower than for the northern hemisphere. 15
i=1 Ai Li
Fig. 7 shows that Group 3 occupies a very narrow strip in LN  = 15
, (17)
the {Tpeak , Qasym } plane, extending from Tpeak ≈ −50 days and i=1 Ai
Qasym ≈ 2.5 to Tpeak ≈ −10 days and Qasym ≈ 6. The place- where 15 latitudinal slabs are considered, for which the cen-
ment of a model nucleus within this stretch is strongly corre- tral latitude is Li and Ai is the active area fraction of a slab
lated with AS /AN ; this ratio decreases from around unity at the (a corresponding quantity LS  is defined for the southern hemi-
lower left end of the distribution, to around AS /AN ≈ 0.5 at the sphere). Note that this definition prevents slabs with a small |Li |
upper right end. (and comparably large surface areas) to get a larger weight than
Furthermore, it is noticeable that Ω  0 cases (black slabs close to the poles.
dots) all are characterized by comparably low Qasym values and Considering model nuclei with  ∈ [−2.5 , −1.5 ],
a rather early production rate peak, while Ω < 0 cases (black Tpeak < −40 days, and AN , AS  30%, we find 25◦  LN  
236 B.J.R. Davidsson, P.J. Gutiérrez / Icarus 180 (2006) 224–242

Fig. 7. Water production rate asymmetry Qasym plotted versus the day of maximum water production rate Tpeak . Here, 103 surface activity maps are considered,
covering a smooth triaxial nucleus with (I, Φ) = (56◦ , 153◦ ) and P = 12.3 h. The LEAM was used in all cases.

64◦ , and −63◦  LS   −21◦ , which must be considered very The three prograde SAOs with P = 12.3 h (Table 3) do not
wide ranges. The conclusion is therefore that nothing certain differ substantially from each other in terms of the upper den-
can be said about the location of active areas on 81P (when sity limit. The only significant differences are the following:
using the LEAM for a smooth triaxial body and the empiri-
cal constraints considered here), not even when these areas are (1) The SAO (I, Φ) = (56◦ , 167◦ ) does not result in any ac-
fairly small. tivity maps yielding a negative Ω when  ∈ [−2.5 ,
−1.5 ] and Tpeak < −40 days.
Group 4 (green symbols) (2) For the SAO (I, Φ) = (75◦ , 150◦ ), nearly half of the model
The fourth group is characterized by a low mass or density, nuclei with  ∈ [−2.5 , −1.5 ] and Tpeak < −40 days
and a  < −2.5 which recedes rapidly from zero as ρbulk have Ω < 0 .
decreases. In Fig. 7, this group consistently has a somewhat (3) For the latter SAO, several activity patterns yield AS /AN ≈
higher production curve asymmetry Qasym for a certain Tpeak , 2, simultaneously fulfilling  ∈ [−2.5 , −1.5 ], Tpeak <
compared with Group 3. However, Fig. 6 shows that Groups 3 −40 days, and even Ω < 0 .
and 4 are very similar in terms of the AS /AN ratio, although
the latter group tend to have a somewhat lower ratio, i.e., an The last observation raises some doubts regarding the pre-
vious statement, that the northern hemisphere possibly has
even stronger difference between the hemispheres in terms of
a larger active area fraction than the southern hemisphere.
ice coverage.
Although the spin axis obliquity hardly could have been as
large as 75◦ during the Stardust flyby (if the spin axis in-
Calculations have also been made for two other prograde deed is aligned with the shortest principal axis), it cannot
spin axis orientations, estimated by Farnham and Schleicher be excluded that I ≈ 75◦ during the 1997 apparition, when
(2005) and Sekanina (2003), plus the three reverse (retrograde) Sekanina (2003) made his analysis, since cometary spin axis
orientations, considering both P = 12.3 h and P = 25 h. Re- orientations may change over time, e.g., due to sublimation-
sults from these simulations are found in Tables 3 and 4. induced torques (e.g., Samarasinha et al., 1996; Jewitt, 1997;
The nucleus bulk density of Comet 81P/Wild 2 237

Table 3
Results obtained for prograde nucleus rotation
(I, Φ) P (h) Body TPM h 1
ρbulk A1N A1S Q1asym 2
ρbulk A2N A2S Q2asym
(56, 153) 12.3 Tri LEAM 0.1 540 22–54 20–49 2.3–3.4 540 24–54 20–47 3.1–3.4
(56, 167) 400 22–54 17–45 1.7–3.0
(75, 150) 500 14–50 27–57 1.7–2.4 500 14–50 27–57 1.7–2.4
(56, 153) 25 490 24–55 21–54 2.1–3.1 390 29–43 28–37 2.7–3.1
(56, 167) 440 20–56 16–47 1.8–2.6
(75, 150) 470 15–48 22–59 1.7–2.4 470 15–48 22–59 1.8–2.4
(56, 153) 12 HECEM 0 400 8–25 13–35 1.5
0.01 430 10–27 14–34 1.6–1.7
0.5 520 9–33 12–33 1.9–2.8 430 28–30 26–27 2.6–2.8
25 530 10–31 13–36 2.0–2.5
12 Irr#1 0 380 7–17 13–26 1.5–1.8
0.01 430 8–20 12–28 1.6–2.0
0.5 500 9–33 6–27 1.8–3.1 500 27–33 18–21 2.8–3.1
25 540 11–34 6–29 2.0–3.1 500 11–34 6–23 2.8–3.1
12 Irr#2 0 430 12–24 17–34 1.5
0.01 460 11–24 16–37 1.6–1.8
0.5 530 15–33 11–30 2.2–3.1 530 20–30 19–30 2.7–2.9
25 580 11–35 14–30 1.8–2.8 460 23–32 23–28 2.5–2.7
12 Irr#3 0 410 9–20 14–32 1.4
0.01 330 10–27 24–28 1.5–1.6
0.5 450 13–30 16–36 1.9–2.4
25 210 16–17 21–24 1.8–2.0
Note. Results obtained for prograde spin axis orientations. Columns 1–5 show the spin axis orientation (SAO), rotational period (P ), nucleus model (Body),
thermophysical model (TPM), and Hertz factor (h), respectively (note that slightly different κd are used for the LEAM and HECEM). Columns 6–9 show results
obtained by requiring  ∈ [−2.5 , −1.5 ] and Tpeak < −40 days; upper limit on the density (ρbulk ) obtained by applying the nucleus volume suggested by
Brownlee et al. (2004) and Duxbury et al. (2004), active area fraction of the northern hemisphere (AN ), active area fraction of the southern hemisphere (AS ), and
the water production curve asymmetry (Qasym ), respectively (superscript “1”). Columns 10–13 show the same four quantities, requiring  ∈ [−2.5 , −1.5 ],
Tpeak < −40 days, and Ω < 0 (superscript “2”).

Table 4
Results obtained for retrograde nucleus rotation
(I, Φ) P (h) Body TPM h 1
ρbulk A1N A1S Q1asym
(124, 333) 12.3 Tri LEAM 0.1 190 53–90 23–37 1.1–1.2
(124, 347)
(105, 330) 310 36–75 22–47 1.4–1.7
(124, 333) 25 290 40–85 15–35 1.1–1.2
(124, 347)
(105, 330) 330 27–71 23–49 1.5–1.7
(124, 333) 12 HECEM 0 400 8–25 13–35 1.5
0.01 300 20–39 11–26 1.4
0.5 290 27–64 8–30 1.1
25 420 25–62 10–33 1.1
12 Irr#1 0 160 20 17 1.4
0.01 460 17–40 7–26 1.4
0.5 130 35 14 1.1
25 120 19–26 9–19 1.1–1.2
12 Irr#2 0 460 14–35 7–20 1.5–1.6
0.01 360 14–39 7–25 1.3–1.5
0.5 290 23–63 8–22 1.0–1.1
25 400 23–63 8–25 1.1
12 Irr#3 0 250 18–26 14–18 1.5–1.7
0.01 310 19–37 10–22 1.5–1.6
0.5 210 33–37 10–18 1.1
25 320 18–53 7–31 1.2–1.3
Note. Results obtained for retrograde spin axis orientations. See caption of Table 3 for details. Notice that the criteria  ∈ [−2.5 , −1.5 ], Tpeak < −40 days,
and Ω < 0 never are fulfilled simultaneously for these models.

Neishtadt et al., 2002; Gutiérrez et al., 2002, 2003; Mysen, A small but noticeable reduction of the estimated upper limit
2004). The AS /AN  0.9 result should therefore be treated with on the bulk density is seen when the spin period is increased to
caution. P = 25 h, demonstrating that ρbulk is somewhat more sensitive
238 B.J.R. Davidsson, P.J. Gutiérrez / Icarus 180 (2006) 224–242

to P than to the spin axis orientation (within the considered (and thereby ρbulk ) in fact is higher in the LEAM, due to layer
{I, Φ} range). The reason for the decrease in ρbulk is a reduc- energy absorption and differences in the thermophysical prop-
tion of the thermal lag angle taking place when the rotational erties, caused by high porosity and the resulting sub-surface
velocity is reduced. sublimation/recondensation and gas diffusion (leading to inter-
Next, the retrograde SAOs are studied (Table 4). In this case, nal transport of both mass and heat). The different, but similar,
it is the southern hemisphere which predominantly is illumi- ρbulk estimates illustrate that the dependence on thermophysi-
nated pre-perihelion, while the northern hemisphere is facing cal model assumptions is not crucial for the ρbulk estimate for
the Sun post-perihelion (e.g., during the Stardust flyby). By the considered models.
reversing the sense of rotation, the non-gravitational force vec- A more substantial difference between LEAM and HECEM
tors change substantially compared to previously studied cases is found when comparing AN and AS . A higher degree of
(they are basically mirrored in the plane defined by the radius ice coverage is predicted by the LEAM (AN and AS are
vector and the SAO). The reason is that the strongest sublima- both 50–60%), than by the HECEM (AN and AS are both
tion always takes place during local afternoon, due to thermal 30–40%). This difference is caused by a more efficient sub-
lag. Substantial differences in the properties and behavior of surface heating in the former model (mainly due to layer energy
the model nuclei compared to prograde rotators are therefore absorption). As a result, the water production rate per surface
expected, which is confirmed by studying Table 4. unit area is lower in the LEAM than in the HECEM, which
means that the former model requires larger fractions of the nu-
(1) The estimated upper limit on the density, within the range cleus to be ice-covered in order to reach a certain total water
440  ρbulk  540 kg m−3 for prograde rotators, decrease production rate.
substantially, to 190  ρbulk  330 kg m−3 for retrograde Due to the smaller active area fraction obtained with the
rotators. HECEM, the investigation regarding LN  and LS  can be re-
(2) The density estimates tend to increase with P , contrary to peated, now with a considerably larger sample of model nuclei.
the behavior of prograde rotators. It is found that the mean latitudes are fairly well constrained
(3) For (I, Φ) = (124◦ , 347◦ ), i.e., the retrograde version of for individual model cases (i.e., a certain combination of shape,
the Farnham and Schleicher (2005) estimate, it is impossi- SAO, P , and h), but that variations from case to case are large.
ble to find surface activity patterns resulting in both  ∈ Considering the whole sample, the ranges 18◦  LN   64◦
[−2.5 , −1.5 ] and Tpeak < −40 days. and −62◦  LS   −13◦ are obtained, again illustrating that
(4) AN reaches very high values in many cases, indicating that the location of active areas on 81P cannot be constrained in the
up to 90% of the northern hemisphere may be covered by present work.
ice. The irregular bodies display a fairly wide range of density
(5) The production rate asymmetry takes comparably low val- estimates, however not dramatically different from that of the
ues, 1.1  Qasym  1.7. smooth triaxial body, ρbulk  210–580 kg m−3 (depending on
(6) In cases where  ∈ [−2.5 , −1.5 ] and Tpeak < −40 days nucleus shape model, P , and h). Although the real shape of
are fulfilled, Ω takes very high positive values. The range 81P has not been studied here, this shows that a large differ-
is 10  Ω  34 for P = 12.3 h and 8  Ω  27 ence between our density estimates and that of the real body is
for P = 25 h, depending on the SAO. None of the empirical not expected, unless the real thermophysical behavior of 81P is
Ω estimates shown in Fig. 1 reach such high values. fundamentally different from that assumed in our models. It is
interesting to note that ρbulk generally increases with h for the
4.2. Smooth and irregular model nuclei studied with the irregular bodies, which indicates that the net thermal lag an-
HECEM gle is not erased by surface topography. It has been speculated
(Samarasinha et al., 1996) that irregular bodies do not display
Tables 3 and 4 also summarize the results obtained by study- net thermal lag even if the conductivity is substantial, due to
ing a smooth triaxial body and three irregular bodies, using the randomization of local surface normals (and thrust vectors),
heat conduction equation model, HECEM. compared to a smooth body. According to our simulations, this
First considering the smooth triaxial model body (prograde does not seem to be the case, at least not under the conditions
rotation), it is interesting to see the effect of an increasing heat studied here [the same conclusion was drawn by Davidsson and
conduction on the estimated upper density limit ρbulk . A body Gutiérrez (2005)].
with no heat conduction (h = 0) yields ρbulk  400 kg m−3 , A final observation for prograde rotators, is that Ω < 0
while h = 0.5 (corresponding to κ ≈ 1.6 W m−1 K−1 at T = cases only appear in the HECEM simulations when the conduc-
200 K) results in a higher density, ρbulk ≈ 520 kg m−3 , due to tivity reaches a certain level, κ  1.6 W m−1 K−1 , regardless of
the larger thermal lag angle. The HECEM h = 0.5 result can the considered P and shape models (except Irr#3, for which no
be compared with the LEAM h = 0.1 (κ ≈ 0.3 W m−1 K−1 Ω < 0 cases are found). If Ω indeed is negative, and if the
at T = 200 K) estimate for the same SAO and P , which is rotation is prograde, this would therefore be a possible observa-
ρbulk  540 kg m−3 . Since κ is a factor 5 smaller for the LEAM tional evidence for a certain degree of heat conductivity in the
compared to the HECEM h = 0.5 case, one could expect that surface material of 81P.
the density estimate should be substantially smaller in the for- The simulations of retrograde rotation shown in Table 4 con-
mer model. The opposite is true, since the thermal lag angle firm the results previously described for the LEAM.
The nucleus bulk density of Comet 81P/Wild 2 239

(1) Retrograde rotation generally implies lower density esti- is much higher, i.e., macroscopic void spaces between con-
mates than prograde rotation, ρbulk  120–460 kg m−3 (de- stituent cometesimals could be present. The results are also
pending on shape, P , and h). consistent with the low or very low bulk density estimates ob-
(2) If only surface activity patterns yielding  ∈ [−2.5 , tained for other comets, or theoretical investigations on come-
−1.5 ] and Tpeak < −40 days are considered, they conse- tesimal growth (Donn, 1963; Rickman, 1986, 1989; Greenberg,
quently result in low water production curve asymmetries, 1986; Samarasinha and Belton, 1995; Asphaug and Benz, 1996;
1.0  Qasym  1.7. Farnham and Cochran, 2002; Davidsson and Gutiérrez, 2004,
(3) In no cases is Ω < 0 obtained. The lowest value found 2005).
is Ω = 3.8 , obtained for the body Irr#2, when h = 0 For retrograde rotators, we consistently find that model nu-
and P = 12 h. Additional cases with Ω  5.0 (remem- clei fulfilling  ∈ [−2.5 , −1.5 ] and Tpeak < −40 days,
bering the highest empirical Ω estimate) are found for always have very low production rate asymmetries, 1.0 
the smooth triaxial body (h = 0, P = 12 h), and Irr#2 Qasym  1.7, and substantial (positive) Ω values, regard-
(h = 0.01, P = 12 h). If only h = 0.5 cases are considered, less of the considered nucleus shape or thermophysical model.
the lower limit ranges from Ω = 9.6 –23.7 (depend- A Qasym close to unity indicates that the water production rate
ing on shape and P ), clearly showing that if the rotation around T = 100 days and beyond should have been nearly as
is retrograde, the conductivity needs to be very low to yield high as during the pre-perihelion peak. Is this scenario at all
consistency even with the highest empirical Ω estimates. likely?
One could argue that the relatively low rates obtained by
5. Discussion Farnham and Schleicher (2005) around T ∼ 60 days post-
perihelion (a factor ∼4 lower than their largest pre-perihelion
5.1. The results measurement), rules out the high post-perihelion activity re-
quired to yield Qasym ≈ 1. However, several of our calculated
Considering all simulations performed here, it is reasonable water production rate curves have a local depression around
to conclude that a likely upper limit on the mass of 81P is T = 50 days, after which the production rate rises substantially,
M  2.3 × 1013 kg. By assuming a nucleus volume as given by yielding a second peak almost as high as the pre-perihelion
Brownlee et al. (2004) and Duxbury et al. (2004), this results in peak (which explains the Qasym ≈ 1 values). This occurs since
the nucleus bulk densities given in Tables 3 and 4, i.e., the upper all these maps have a comparably high AN . The local mini-
limit is ρbulk  600 kg m−3 . The slightly smaller nucleus pro- mum in Qcalc is produced when the sub-solar point reaches
posed by Howington-Kraus et al. (2005) leads to a somewhat the equator (minimization of the mean illuminated cross sec-
higher upper limit on the bulk density, ρbulk  800 kg m−3 . We tion), and the second peak forms as the sub-solar point starts
have tried to minimize the model assumption dependence of reaching the ice-rich northern hemisphere (again reminding that
these estimates by considering a variety of relevant model nu- retrograde rotation is assumed). Simulations performed with
cleus shape types, spin axis orientations, rotational periods, and the LEAM show that test nuclei with a substantially lower AN ,
thermophysical properties (i.e., in terms of the porosity, trans- which would yield a much larger Qasym (i.e., comparably low
parency to optical radiation, and conductivity of the surface production rates beyond T = 60 days), apparently have been
material). removed by the ρbulk > 100 kg m−3 criterion (and not by the
The density of fully compacted comet material is likely to −2.5    −1.5 criterion). Is it therefore possible that
be in the range 1200  ρcompact  1650 kg m−3 . The lower the empirical water production rate estimates, which do not ex-
limit is obtained by assuming a 0.2:0.19:0.55:0.06 mixture (by tend beyond T ≈ 60 days, captured a broad local minimum in
mass) of silicates, organics, volatiles, and carbon (Greenberg the production rate, but fail to reveal a sub-sequent intensifica-
and Hage, 1990), and the following densities for these com- tion of the water production, i.e., that Qasym ≈ 1 could be a real
ponents; ρsilicate = 2950 kg m−3 (as for a mixture of 70% possibility?
pyroxene, here MgSiO3 , and 30% olivine, here MgFeSiO4 ), Perhaps a clue to this problem is found in coma morphology
ρorganics = 1600 kg m−3 , ρvolatiles = 933 kg m−3 (assuming pure observations, which extend further in time than the water pro-
water ice), and ρcarbon = 2250 kg m−3 (as for graphite). The duction estimates. In July 1997 (T = 63 days), Farnham and
upper limit was proposed by Greenberg (1998), using slightly Schleicher (2005) made a clear detection of a coma jet, indi-
different mass fractions, and generally larger densities for the cating that the water production was fierce enough at the time
constituent materials. to carry away substantial amounts of dust. In October the same
Combining the previously mentioned ranges for ρbulk and year (T = 148 days), no well-defined coma features could be
ρcompact , this implies a lower limit on the nucleus bulk poros- seen. At an intermediate (post-perihelion) position in the or-
ity ψ of 0.3–0.6. The lower value is consistent with random bit, at T = 101 days (but during the 2003/2004 apparition),
close packing of spheres (e.g., of the sub-µm grains which the coma showed no signs of jet structures either. The time se-
presumably make up a comet nucleus). The higher value is quence T = 63, 101, 148 days, corresponds to a change in the
consistent with planetesimal growth by gradual agglomera- sub-solar latitude of LN = 8◦ , 23◦ , 37◦ (remembering that ret-
tion of small fractal grain clusters, including a mild colli- rograde rotation is discussed here), i.e., the gradual switch-off
sion compaction (e.g., Blum, 2004). Note that the range in of comet activity hinted by the disappearance of the jet, appears
ψ denotes a lower limit, i.e., it cannot be excluded that ψ to support a comparably low AN . If the water production rate in-
240 B.J.R. Davidsson, P.J. Gutiérrez / Icarus 180 (2006) 224–242

deed had increased a factor ∼3–4 between T ≈ 60 to 150 days, possibility is that a thin porous dust mantle covers the entire
it is somewhat surprising that this did not result in a rejuvena- nucleus, through which vapor diffuses at a quenched rate. We
tion of the coma, e.g., in form of reappearing jet structures. note that the existence of near-nucleus coma jet structures, as
Another indication that Qasym ≈ 1 may be unreasonable observed by Stardust, do not necessarily rule out surface homo-
comes from the visual lightcurve from the 1990 apparition, geneity of these kinds, since such features, in principle, can be
which covers the period −60  T  230 days (see, e.g., formed by surface topography (e.g., Crifo and Rodionov, 1997;
Sekanina, 2003). The intrinsic brightness peaks about three Crifo et al., 2003).
weeks before perihelion, and the subsequent fading is rapid To crudely test the effect of large-scale chemical homogene-
at first, but slows down substantially around T ≈ 100 days, af- ity of the surface layer (i.e., not necessarily the entire volume
ter which the brightness remains nearly constant for the next of the nucleus), the LEAM is applied for the smooth triax-
∼100 days (incidentally, we note the qualitative similarity be- ial body with (I, Φ) = (56◦ , 153◦ ) and P = 12.3 h, assuming
tween the shapes of the lightcurve and the calculated mean that all surface facets are potentially active, but reducing the
water production rate in Fig. 4—the latter changes rather lit- local water production rate (and non-gravitational force mag-
tle in the T ≈ 100–200 day region, since the diminishing solar nitude) a factor fscale for each facet on a trial nucleus. From
flux is compensated by an increasing mean illuminated cross a large sample of trial nuclei (with 0.01  fscale  1), 103
section). The lightcurve peak and post-perihelion plateau dif- cases are selected, requiring that the total water production rate
fer about two magnitudes in brightness, which may exclude a curves are confined to the envelope shown in Fig. 2, and that
second high peak in the water production rate, although the ρbulk > 100 kg m−3 . Model nuclei fulfilling these criteria have
correlation between coma brightness and water production rate 0.15  fscale  0.52, again illustrating that a fully active model
certainly is not clear (keeping in mind the possibility that the nucleus (i.e., fscale = 1) is an unlikely representation of 81P.
dust-to-gas mass ratio, dust size distribution, dust albedo, and For the nominal P [and applying the nucleus volume con-
phase angle may vary substantially over time). sistent with the shape model by Brownlee et al. (2004) and
Although a production curve asymmetry as low as 1.0  Duxbury et al. (2004)], the resulting density estimates have
Qasym  1.7 cannot be excluded for the time being, we con- the range 100 < ρbulk  540 kg m−3 , while  ≈ −1.3 and
sider it rather unlikely. The high positive Ω obtained for
Ω ≈ 1.1 for all test nuclei. The invariant  and Ω can
retrograde rotators (in particular for bodies with at least some
be understood by inserting Eq. (1), solved for M, into Eqs. (2)
degree of conductivity), constitutes an additional, possibly se-
and (3)—the latter equations then only depend on the functional
rious, problem. We therefore consider a retrograde rotation of
form of the non-gravitational force F̄ , not on the absolute value
81P inconsistent with the empirical constraints used in this
|F̄ | (and therefore not on fscale ). The crescent-shaped distribu-
work, provided that our models are sufficiently accurate. We
tion seen in Fig. 5 is therefore caused by differences in the func-
note that this is contrary to the result obtained by Sekanina et
tional form of the non-gravitational force (reflecting differences
al. (2004), but that Farnham and Schleicher (2005) also favor a
in the surface activity patterns), whereas the homogeneous nu-
prograde rotation.
clei would show up on a horizontal line in that figure. By adjust-
5.2. The model assumptions ing P modestly upwards (still within its error bars, as defined
in Section 2.2), the  -value for the homogeneous nuclei can
It is important to discuss a number of assumptions made in be pushed into the empirical range ( ∈ [−2.5 , −1.5 ]).
our modeling, which potentially could limit the validity of our This manoeuvre is accompanied by a slight reduction of the
results. Perhaps the most important of these are the assumption density range, 100 < ρbulk  470 kg m−3 . It is therefore possi-
that active areas exist (i.e., large-scale chemical surface hetero- ble for a model nucleus, lacking inactive regions but having a
geneity), and that sublimation mainly takes place from exposed certain degree of overall water production quenching, to simul-
surface ice (or from a near-surface layer). taneously fulfill the empirical P and  criteria, and having
All nucleus models considered here, which successfully re- its water production curve within the envelope. As for a fully
produce the empirical criteria, indicate that the gas production active model nucleus, all selected test nuclei yield Qasym ≈ 3.8
of 81P is substantially lower than expected from a body uni- and Tpeak ≈ −45 days. In principle, a homogeneous model nu-
formly covered by a thermally coupled ice/dust mixture (illus- cleus therefore fulfills the considered empirical criteria just as
trated by AN , AS  60% or 40% for the LEAM and HECEM, well as a model nucleus with active and inactive regions, and
respectively). This apparently necessary reduction of the water there is no possibility to favor one of these models as a supe-
production rate is here achieved by postulating that certain areas rior representation of 81P. One exception concerns Ω—if this
on 81P are inactive, i.e., do not emit vapor under any illumina- quantity indeed is negative, then surface heterogeneity is re-
tion condition. However, there are alternative ways to lower the quired, since homogeneous nuclei exclusively yield Ω > 0 .
water production rate, also for surfaces which are globally ho- However, and most importantly, the nucleus bulk density esti-
mogeneous, e.g., if there is no thermal coupling between ice and mate is very similar for homogeneous and heterogeneous model
dust (Crifo, 1997). In such a case, large dust grains are much bodies, i.e., ρbulk  600 (or 800) kg m−3 is indeed a plausi-
hotter than the surrounding ice, and substantial amounts of en- ble limit. Additional tests have also been made for the irregular
ergy is reradiated to space, which lowers the water production model nuclei, using a quenched HECEM, yielding very similar
rate per unit area (for a certain illumination condition). Another results.
The nucleus bulk density of Comet 81P/Wild 2 241

It should also be noted, that if the retrograde SAO (I, Φ) = (2) The peak in the water production curve should take place
(124◦ , 333◦ ) is studied, then homogeneous model nuclei (as- 40 days before perihelion, or earlier.
suming a quenched LEAM and a smooth triaxial nucleus shape (3) The nucleus bulk density (calculated by applying the em-
or a quenched HECEM for any of the bodies in this study) fail pirical P) should be in excess of 100 kg m−3 .
to yield P > 0 days, which must be considered a particularly (4) The change in the longitude of perihelion should be within
strong empirical criterion. The plausibility of a retrograde ro- the range −2.5    −1.5 .
tation, which is weak for heterogeneous bodies, therefore gets
even weaker if surface homogeneity is assumed. Furthermore, test nuclei resulting in a water production rate
A second objection that may be raised against our models, during the period 100  T  200 days which is nearly as high
is that sublimation of surface ice (or of a near-surface layer) as during the period −100  T  0 days, and/or having a
perhaps is not the principle outgassing mechanism. Yelle et al. change in the longitude of the ascending node which is sub-
(2004) have suggested that sub-surface voids with sublimat- stantially larger than Ω ∼ 5 , have been considered unlikely
ing walls produce most of the vapor, which exudes through representations of 81P. Using these conditions, the results can
nozzle-like channels or cracks in an overlying dust mantle, pos- be summarized as follows:
sibly exerting a substantial thrust force on the nucleus. In such
a case, the non-gravitational force could be stronger than as- (1) A likely upper limit on the nucleus mass is M  2.3 ×
sumed in the present work, which means that our suggested 1013 kg, which leads to an upper bulk density limit of
density may be an underestimate. However, since local thrust ρbulk  600 − 800 kg m−3 (all values reported in Tables 3
vectors can have virtually any orientation with respect to the and 4 are smaller than this, but the value depends on the ap-
local surface normal, depending on the detailed geometry of plied nucleus volume). The porosity limit is ψ  0.3–0.6.
the nozzle axes, it is not certain that a such a “geyser mod- (2) The nucleus most probably has a prograde rotation, im-
el” indeed would yield a higher density estimate than obtained plying that Stardust primarily imaged the southern hemi-
here. We strongly encourage colleagues in the modeling com- sphere, and that the SAO (i.e., the positive rotational pole)
munity to investigate the properties of such vents, e.g., by at the time was close to (I, Φ) = (56◦ , 153◦ ).
studying coupled thermophysical and gas kinetic models, to (3) The total ice coverage of the nucleus surface may be as
see if or when the sub-surface temperature, vapor production low as 6% and as high as 60%. This is admittedly a poor
rate, and gas pressure reach levels at which vent thrusting be- constraint, caused by the size of the considered water pro-
comes significantly stronger than surface sublimation thrust- duction envelope. The degree of ice coverage is probably
ing. similar between the hemispheres, although the northern
We finally note that it also has been assumed that the spin part of the nucleus may be somewhat more ice-rich than
axis orientation does not change substantially during a perihe- the southern, if the spin axis obliquity indeed was I ∼ 50◦
lion passage. Since similar density estimates were obtained for also in 1997.
a number of SAOs separated by 20◦ , we do not expect this (4) If Ω is negative, as many but not all empirical estimates
assumption to be critical. As mentioned previously, the rota- indicate, our modeling suggests that surface heterogeneity
tional state of 81P and its possible evolution due to sublimation- (active versus inactive areas) most likely is present on 81P,
induced torques will be studied in a forthcoming paper. and that a certain degree of heat conduction of the surface
material is necessary (a Hertz factor h > 0.01).
6. Summary (5) Smooth and irregular model nuclei with comparable di-
mensions and thermophysical properties result in similar
The purpose of this work was to constrain a number of non-gravitational force vectors (leading to similar ρbulk es-
nucleus parameters for Comet 81P/Wild 2, primarily its nu- timates for the same P). This also implies that large-scale
cleus bulk density, by considering a few smooth and irregular surface topography does not erase the net thermal lag (at
model nuclei, having relevant dimensions, rotational periods, least not for the shapes studied here).
(prograde and retrograde) spin axis orientations, and displaying
different thermophysical behavior.
Assuming that the nucleus surface consists of both ice-rich Acknowledgments
and -poor regions, more than 4 × 104 surface activity patterns
have been studied for these model nuclei, in each case calcu- We are very grateful to Steven Chesley and Sławomira Szu-
lating the water production rate and non-gravitational force as towicz for having provided us with several of their computa-
functions of time, the nucleus bulk density required to repro- tions of non-gravitational parameters for Comet 81P/Wild 2.
duce the empirical P, and the values of  and Ω. Four We also thank Teemu Mäkinen for sharing his re-calibrated wa-
conditions have been imposed before a model nucleus qualifies ter production rate data. The comments and suggestions made
as a possible representation of 81P. by Tony Farnham and an anonymous referee were highly appre-
ciated, and helped us improving the presentation of the material.
(1) The water production curve should be within the envelope Both authors acknowledge financial support from the European
shown in Fig. 2 at all times. Space Agency (ESA).
242 B.J.R. Davidsson, P.J. Gutiérrez / Icarus 180 (2006) 224–242

References Gutiérrez, P.J., Ortiz, J.L., Rodrigo, R., López-Moreno, J.J., Jorda, L., 2002.
Evolution of the rotational state of irregular cometary nuclei. Earth Moon
Asphaug, E., Benz, W., 1996. Size, density, and structure of Comet Shoemaker– Planets 90 (1), 239–247.
Levy 9 inferred from the physics of tidal breakup. Icarus 121, 225–248. Gutiérrez, P.J., Jorda, L., Ortiz, J.L., Rodrigo, R., 2003. Long-term simulations
Blum, J., 2004. Grain growth and coagulation. In: Witt, A., Clayton, G., Draine, of the rotational state of small irregular cometary nuclei. Astron. Astro-
B. (Eds.), Astrophysics of Dust. In: ASP Conf. Ser., vol. 309. The Astro- phys. 406, 1123–1133.
nomical Society of the Pacific, San Francisco, pp. 369–391. Howington-Kraus, E., Kirk, R.L., Duxbury, T.C., Hörz, F., Brownlee, D.E.,
Brownlee, D.E., 11 colleagues, 2004. Surface of young Jupiter family Comet Newburn, R.L., Tsou, P., The Stardust Team, 2005. Topography of the
81P/Wild 2: View from the Stardust spacecraft. Science 304, 1764–1769. 81P/Wild 2 nucleus from Stardust stereoimages. Asia–Oceania Geosciences
Crifo, J.F., 1997. The correct evaluation of the sublimation rate of dusty ices Society 2nd Annual Meeting, Singapore. Poster presentation, 58-PS-
under solar illumination, and its implication on the properties of P/Halley A0956.
nucleus. Icarus 130, 549–551. Jewitt, D., 1997. Cometary rotation: An overview. Earth Moon Planets 79 (1),
Crifo, J.F., Rodionov, A.V., 1997. The dependence of the circumnuclear coma 35–53.
structure on the properties of the nucleus. II. First investigation of the coma Kissel, J., Krueger, F.R., Silén, J., Clark, B.C., 2004. The cometary and inter-
surrounding a homogeneous, aspherical nucleus. Icarus 129, 72–93. stellar dust analyzer at Comet 81P/Wild 2. Science 304, 1774–1776.
Crifo, J.F., Loukianov, G.A., Rodionov, A.V., Zakharov, V.V., 2003. Navier–
Mäkinen, J.T.T., Silén, J., Schmidt, W., Kyrölä, E., Summanen, T., Bertaux,
Stokes and direct Monte Carlo simulations of the cicumnuclear coma.
J.-L., Quémerais, E., Lallement, R., 2001. Water production of Comets
II. Homogeneous, aspherical sources. Icarus 163, 479–503.
2P/Encke and 81P/Wild 2 derived from SWAN observations during the
Crovisier, J., Colom, P., Gérard, E., Bockelée-Morvan, D., Bourgois, G., 2002.
1997 apparition. Icarus 152, 268–274.
Observations at Nançay of the OH 18-cm lines in comets. The data base.
Marsden, B.G., Sekanina, Z., Yeomans, D.K., 1973. Comets and nongravita-
Observations made from 1982 to 1999. Astron. Astrophys. 393, 1053–1064.
tional forces. V. Astron. J. 78 (2), 211–225.
Davidsson, B.J.R., 2001. Tidal splitting and rotational breakup of solid biaxial
ellipsoids. Icarus 149, 375–383. Mysen, E., 2004. Rotational dynamics of subsolar sublimating triaxial comets.
Davidsson, B.J.R., Gutiérrez, P.J., 2004. Estimating the nucleus density of Planet. Space Sci. 52 (10), 897–907.
Comet 19P/Borrelly. Icarus 168 (2), 392–408. Neishtadt, A.I., Scheeres, D., Sidorenko, V.V., Vasiliev, A.A., 2002. The influ-
Davidsson, B.J.R., Gutiérrez, P.J., 2005. Nucleus properties of Comet ence of reactive torques on comet nucleus rotation. Icarus 157 (1), 205–218.
67P/Churyumov–Gerasimenko estimated from non-gravitational force Rickman, H., 1986. Masses and densities of Comets Halley and Kopff. In:
modeling. Icarus 176 (2), 453–477. Melita, O. (Ed.), The Comet Nucleus Sample Return Mission. ESA Pub-
Davidsson, B.J.R., Skorov, Y.V., 2002a. On the light-absorbing surface layer of lications Division, ESTEC, Noordwijk, pp. 195–205.
cometary nuclei. I. Radiative transfer. Icarus 156, 223–248. Rickman, H., 1989. The nucleus of Comet Halley: Surface structure, mean den-
Davidsson, B.J.R., Skorov, Y.V., 2002b. On the light-absorbing surface layer of sity, gas and dust production. In: Gombosi, T.I., Atreya, S.K., Grün, E., Han-
cometary nuclei. II. Thermal modeling. Icarus 159, 239–258. ner, M.S. (Eds.), Cometary Environments. Adv. Space Res. 9 (3), 59–71.
Davidsson, B.J.R., Skorov, Y.V., 2004. A practical tool for simulating the Samarasinha, N.H., Belton, M.J.S., 1995. Long-term evolution of rota-
presence of gas comae in thermophysical modeling of cometary nuclei. tional states and nongravitational effects for Halley-like cometary nuclei.
Icarus 168 (1), 163–185. Icarus 116, 340–358.
Donn, B., 1963. The origin and structure of icy cometary nuclei. Icarus 2, 396– Samarasinha, N.H., Mueller, B.E.A., Belton, M.J.S., 1996. Comments on the
402. rotational state and non-gravitational forces of Comet 46P/Wirtanen. Planet.
Duxbury, T.C., Newburn, R.L., Brownlee, D.E., 2004. Comet 81P/Wild 2 size, Space Sci. 44 (3), 275–281.
shape, and orientation. J. Geophys. Res. 109 (E12S02), 1–4. Sekanina, Z., 1981. Rotation and precession of cometary nuclei. Ann. Rev.
Farnham, T.L., Cochran, A.L., 2002. A McDonald Observatory study of Comet Earth Planet. Sci. 9, 113–145.
19P/Borrelly: Placing the Deep Space 1 observations into a broader context.
Sekanina, Z., 1993. Effects of discrete-source outgassing on motions of periodic
Icarus 160 (2), 398–418.
comets and discontinuous orbital anomalies. Astron. J. 105 (2), 702–735.
Farnham, T.L., Schleicher, D.G., 2005. Physical and compositional studies of
Sekanina, Z., 2003. A model for Comet 81P/Wild 2. J. Geophys. Res. 108
Comet 81P/Wild 2 at multiple apparitions. Icarus 173 (2), 533–558.
(E10). SRD2-1-SRD2-14.
Fink, U., Hicks, M.P., Fevig, R.A., 1999. Production rates for the Stardust mis-
sion target: 81P/Wild 2. Icarus 141, 331–340. Sekanina, Z., Brownlee, D.E., Economou, T.E., Tuzzolino, A.J., Green, S.F.,
Greenberg, J.M., 1986. Fluffy comets. In: Lagerkvist, C.-I., Lindblad, B.A., 2004. Modeling the nucleus and jets of Comet 81P/Wild 2 based on the
Lundstedt, H., Rickman, H. (Eds.), Asteroids, Comets, Meteors II. Uppsala Stardust encounter data. Science 304, 1769–1774.
University, Uppsala, pp. 221–223. Tuzzolino, A.J., Economou, T.E., Clark, B.C., Tsou, P., Brownlee, D.E., Green,
Greenberg, J.M., 1998. Making a comet nucleus. Astron. Astrophys. 330, 375– S.F., McDonnell, J.A.M., McBride, N., Colwell, M.T.S.H., 2004. Dust mea-
380. surements in the coma of Comet 81P/Wild 2 by the Dust Flux Monitor
Greenberg, J.M., Hage, J.I., 1990. From interstellar dust to comets: A unifica- Instrument. Science 304, 1776–1780.
tion of observational constraints. Astrophys. J. 361, 260–274. Yelle, R.V., Soderblom, L.A., Jokipii, J.R., 2004. Formation of jets in Comet
Gutiérrez, P.J., Ortiz, J.L., Rodrigo, R., López-Moreno, J.J., 2001. Effects of 19P/Borrelly by subsurface geysers. Icarus 167, 30–36.
irregular shape and topography in thermophysical models of heterogeneous Yeomans, D.K., Chodas, P.W., 1989. An asymmetric outgassing model for
cometary nuclei. Astron. Astrophys. 374, 326–336. cometary nongravitational accelerations. Astron. J. 98 (3), 1083–1093.

You might also like