Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Icarus 160, 300–312 (2002)

doi:10.1006/icar.2002.6976

Long-Term Evolution of Objects in the Kuiper Belt Zone—Effects


of Insolation and Radiogenic Heating
Young-Jun Choi, Merav Cohen, Rainer Merk, and Dina Prialnik
Department of Geophysics and Planetary Sciences, Tel Aviv University, Ramat Aviv 69978, Israel
E-mail: dina@planet.tau.ac.il

Received November 11, 2001; revised July 21, 2002

spectrum of a Kuiper Belt object (KBO), indicating a kinship


The Kuiper Belt zone is unique insofar as the major heat sources to comets (Brown et al. 1999). In fact, it has long been sug-
of objects a few tens of kilometers in size—solar radiation on the gested by Duncan et al. (1988) and others that the Kuiper Belt is
one hand and radioactive decay on the other—have comparable the source of short-period comets and of Centaurs (small bod-
power. This leads to unique evolutionary patterns, with heat waves ies characterized by dynamically unstable orbits between Jupiter
propagating inward from the irradiated surface and outward from and Neptune). The latter are probably transition objects from the
the radioactively heated interior. A major radioactive source that is Kuiper Belt reservoir to the inner Solar System (Gladman and
considered in this study is 26 Al. The long-term evolution of several
Duncan 1990). Their nature—whether asteroidal or cometary—
models with characteristics typical of Kuiper Belt objects is followed
is uncertain; they show evidence of complex organic molecules,
by means of a 1-D numerical code that solves the heat and mass
balance equations on a spherically symmetric grid. The free pa- and water ice was detected on the surfaces of several such objects
rameters considered are radius (10–500 km), heliocentric distance (Brown and Koresko 1998, Brown et al. 1998, Luu et al. 2000).
(30–120 AU), and initial 26 Al content (0–5 × 10−8 by mass). The ini- Observed Kuiper Belt objects are larger than typical comets,
tial composition assumed is a porous mixture of ices (H2 O, CO, and their radii ranging between a few tens and a few hundred of
CO2 ) and dust. Gases released in the interior are allowed to escape to kilometers (Jewitt et al. 1998, Gladman et al. 1998, Chiang and
the surface. It is shown that, depending on parameters, the interior Brown 1999, Sheppard et al. 2000, Trujillo et al. 2001). The
may reach quite high temperatures (up to 180 K). The models sug- largest objects have radii of about 500 km (e.g., Varuna, 450 km,
gest that Kuiper Belt objects are likely to lose the ices of very volatile according to Jewitt et al. 2001).
species during early evolution; ices of less volatile species are re- Whether KBOs are pristine, or else, have been altered ei-
tained in a surface layer, about 1 km thick. The models indicate that
ther thermally or in terms of structure, is still an open question.
the amorphous ice crystallizes in the interior, and hence some objects
Their sizes may have been significantly altered by collisions
may also lose part of the volatiles trapped in amorphous ice. Gen-
erally, the outer layers are far less affected than the inner part, re- (Davis and Farinella 1997), but the low collision velocities in the
sulting in a stratified composition and altered porosity distribution. Kuiper Belt could not significantly affect their thermal structure
These changes in structure and composition should have significant (Orosei et al. 2001). Thus the main heat sources that could affect
consequences for the short-period comets, which are believed to be them are insolation and radioactive decay, which determine the
descendants of Kuiper Belt objects. c 2002 Elsevier Science (USA) evolution of comets. In the region of the inner Solar System, the
Key Words: comets; comets, composition; ices; Kuiper Belt ob- Sun’s radiation is the dominant heat source, and the outer layers
jects. of a comet nucleus are those most affected by it. Far from the
Sun, where comets spend most of their lifetimes, internal heat-
ing by radionuclides may play an important role. Besides 40 K,
235
1. INTRODUCTION U, 238 U, and 232 Th, which constitute the major heat sources
in terrestrial planets, the radioactive isotope 26 Al has long been
The region of the Solar System beyond Neptune’s orbit, known recognized as a potential heat source capable of melting bodies
as the Kuiper (or Edgeworth–Kuiper) Belt, is populated by small with radii between 100 and 1000 km (Urey 1955). Its lifetime
bodies, which represent the surviving remnants of the once (half-life of 7.4 × 105 years) is just long enough for it to out-
more massive protoplanetary disk of planetesimals. Having been live the formation of comets and small icy satellites, and short
formed farthest out from the Sun, they are believed to consti- enough to generate sufficient heat power for overcoming cool-
tute the most primitive, least thermally processed matter. Since ing by conduction. Although radioactive heating is expected to
1992, when the first object in this class was discovered, over 560 have affected comets to a much lesser extent than satellites and
more have been detected, with semi-major axes between 35 and planets, it was found to cause an internal temperature rise of
over 100 AU. Recently, evidence of H2 O was discovered in the tens of kelvins (e.g., Whipple and Stefanic 1966, Wallis 1980,
300

0019-1035/02 $35.00
c 2002 Elsevier Science (USA)
All rights reserved.
THERMAL EVOLUTION IN THE KUIPER BELT 301

TABLE I centric distance dH , radius R, and the abundance of radioactive


Radioactive Isotope Characteristics species X rad
Isotope τ (yr) X0 H (erg/g) Q 0 (erg/g/yr) 1 R2 4π 3 X rad Hrad
(1 − A)L  2 = R ρ , (1)
40 K 1.82(9) 1.1(−6) 1.72(16) 10.7 4 dH 3 τrad
232 Th 2.0(10) 5.5(−8) 1.65(17) 0.454
238 U 6.50(9) 2.2(−8) 1.92(17) 0.645 where A is the albedo, L  is the solar luminosity, τrad is the char-
235 U
26 Al
1.03(9) 6.3(−9) 1.86(17) 1.13 acteristic radioactive decay time, and Hrad is the energy released
1.06(6) 6.7(−7) 1.48(17) 9.4(4)
per unit mass. If we consider only 26 Al, which is the most pow-
erful source, adopt A = 0.04, ρ = 0.7 g cm−3 , and normalize the
parameters,
Consolmagno and Lewis 1978, Schubert et al. 1981, Elsworth
and Schubert 1983, Prialnik et al. 1987, Prialnik and Bar-Nun x = X rad / X rad,0 X rad,0 = 6.7 × 10−7
1990, Haruyama et al. 1993, Yabushita 1993, Prialnik and y = dH /dH,0 dH,0 = 30 AU
Podolak 1995, and Podolak and Prialnik 1997). Moreover, when
very low thermal conductivities are assumed, a runaway increase z = R/R0 R0 = 10 km,
in the internal temperature occurs and most of the water ice
crystallizes (if it is initially amorphous), although for a suffi- we obtain the following relation in three variables
ciently large conductivity, the amorphous ice may be almost
completely preserved. Typical abundances—mass fractions of x y 2 z = 0.55, (2)
the dust component—of radioactive isotopes in chondritic me-
teorites (MacDonald 1959), and other relevant characteristics, which is illustrated by the shaded surface in Fig. 1. Above this
are given in Table I. surface radioactive heating dominates, while solar heating dom-
It is instructive to compare the power supplied by these two inates below it. The interesting conclusion of this exercise is that
major energy sources, in terms of the leading parameters: helio- the Kuiper Belt zone is quite unique within the Solar System in

x y2z=0.55

50

45

40

35

30

25
z

20

15

10

5 1
2
0
0 3
−0.5
−1 4
−1.5
−2 5
y

log (x)
10

FIG. 1. Solution of Eq. (2), showing the plane where solar heating and radiogenic energy release have equal power, in the distance—radius—26 Al-content
space.
302 CHOI ET AL.

the sense that the major heat sources—solar radiation on the Wetherill and Stewart 1989). Thus, the comparably short life-
one hand and radioactive decay on the other—have compara- time of 26 Al and the growth time of asteroidal bodies are not
ble power for objects of the observed KBO size. This should mutually contradictory (Ghosh and McSween 1998). Further-
lead to unique evolutionary patterns. When radioactive heat- more, the decay product 26 Mg can be found experimentally in
ing dominates (e.g., in the Oort cloud), evolution (or change Ca–Al inclusions of meteorites, the best-known among them be-
in structure and composition) proceeds from the center out- ing Allende (Lee et al. 1976). Srinivasan et al. (1999) detected
ward. By contrast, in the inner Solar System, where insolation for the first time 26 Mg in a differentiated meteorite and thus could
dominates, structural changes propagate from the surface in- confirm the role of 26 Al in the differentiation of meteoritic parent
ward. Kuiper Belt objects may experience both, and the resulting bodies. Renewed interest in this radionuclide followed the de-
structure may be quite unexpected. In this context, we should tection of interstellar 1.809 Mev γ -rays from the decay of 26 Al
keep in mind that although the two energy sources are compa- (Mahoney et al. 1984, Share et al. 1985, Clayton and Leising
rable in power, they differ in character: radioactivity is a body 1987).
source, homogeneously distributed, and time-dependent; solar Accretion models for KBOs (Kenyon and Luu 1998, Farinella
radiation is a surface source and, for KBOs which have close et al. 2000) or for cometary icy bodies (Weidenschilling 1997)
to circular orbits, it is almost constant in time, although its ef- show that accretion times are in the range between 105 years
fect should decrease as the surface temperature reaches equi- (Weidenschilling 1997) and several million years (Kenyon and
librium. Luu 1998, Farinella et al. 2000). An upper limit for accretion
The thermal evolution of distant comets (dH = 100 AU) under time scales in the Kuiper Belt region seems to be the forma-
the effect of radiogenic heating, with special emphasis on 26 Al, tion time of Neptune, since it is assumed that the formation of
was studied by Prialnik and Podolak (1999) for a range of radii Neptune efficiently terminated growth in the Kuiper Belt re-
and porosities, but only for relatively short time spans—until the gion (Farinella et al. 2000). Kenyon and Luu (1998) assumed
rising central temperature attained maximum. Only water ice a value of 50–100 Myr (Pollack et al. 1996) as an upper limit
was considered in that study, and volatiles trapped in the amor- of this time scale. Models based on Wetherill’s particle-in-a-
phous ice. Changes in structure and composition thus resulted box method (Wetherill and Stewart 1989), such as that used by
only when the temperatures were sufficiently high for crystal- Kenyon and Luu (1998), lead to longer accretion times than
lization and gas release to occur. In those cases a layered struc- models that include the transition between drag-dominated and
ture emerged, with ices of more volatile species overlying ices gravity-dominated regimes (Weidenschilling 1997). The former
of less volatile species. Recently, the thermal evolution of KBOs models start with a seed body and exclude sticking effects,
was considered by De Sanctis et al. (2001), who evolved models whereas the latter models describe the growth of particles from
of two such objects for up to 10 Myr, including the long-lived micrometer to kilometer size. Weidenschilling (1997) showed
radioactive species and, in one case, 26 Al as well, and assuming a that it is possible to grow large icy bodies of a radius of about
composition of mixed ices and dust. They found that the surface 40 km within 2.5 × 105 years in the region at 30 AU. These in-
layers became depleted of the most volatile species. However, vestigations show that the short-lived isotope 26 Al can be consid-
the evolution was not followed long enough for the effect of ered an important contributor to internal heat generation within
radiogenic heating to become fully developed and mainly the KBOs.
effect of insolation was, in fact, examined. In the present study All the observational evidence points toward an interstellar
we follow, numerically, the long-term evolution of KBO models isotopic ratio of 26 Al/ 27Al ≈ 5 × 10−5 , implying an initial mass
over the [R, dH , X 0 (26 Al)] parameter space. The 26 Al issue is fraction X 0 (26 Al) ≈ 7 × 10−7 in dust (rock) and presumably an
further discussed in the next section. The model, assumptions, order of magnitude less in objects whose time of aggregation
and parameters are addressed in Section 3; the results of nu- did not exceed a few million years. Therefore, an initial 26 Al
merical computations are described and discussed in Section 4, mass fraction of a few 10−8 should be a reasonable working
focusing on volatile depletion and loss of homogeneity in struc- assumption.
ture as well as in composition. Finally, our main conclusions are
summarized and compared to other studies in Section 5.
3. MODEL, ASSUMPTIONS, AND PARAMETERS

2. 26
AL AND THE FORMATION OF SMALL BODIES We use a numerical code that solves the equations of mass
IN THE SOLAR SYSTEM and energy conservation for a spherically symmetric porous nu-
cleus. The nucleus is assumed to be composed of dust and a
The radionuclide 26 Al is already widely used in thermal mod- mixture of volatiles, which may be found in a solid or gaseous
eling of asteroidal bodies (Miyamoto et al. 1981, Akridge et al. state. The dust is assumed to include radioactive elements in
1998, Ghosh and McSween 1998). For silicate small bodies such abundances typical of meteorites. The water ice is assumed to
as asteroids, accretion models predict growth times in the range be initially amorphous, but crystallization is taken into account
between 104 years and 1 Myr, depending on whether or not according to the temperature-dependent rate λ(T ) = 1.05 × 1013
conditions allow for runaway accretion (Weidenschilling 1988, exp (−5370/T ) s−1 , given by Schmitt et al. (1989). In some test
THERMAL EVOLUTION IN THE KUIPER BELT 303

cases we also consider gases that are trapped in the amorphous The system of non-linear partial differential equations is solved
ice and released upon crystallization. numerically on a one-dimensional spherical grid. The radial res-
The equations of mass conservation for a volatile species α in olution is higher near the surface, where gradients are steeper.
gaseous (subscript g) and solid (subscript s) phases are, respec- The time steps are adjusted by the code to keep the changes
tively, in temperature during one time step confined to a few percent.
Further details on the modeling of the porous medium, and on
∂ρα,g the method of computing the sublimation rate, may be found in
+ ∇ · Jα = q α , (3)
∂t Mekler et al. (1990); details of the implicit, iterative numerical
∂ρα,s code are given by Prialnik (1992).
= −qα , (4) The working assumption of the present calculations is that
∂t
gases released in the interior can easily escape, that is, can
where ρ, with the appropriate subscript, denotes the bulk density channel easily to the surface and out of the nucleus. This is
(mass per unit volume of cometary material) of a species in a justified by the very low strength of cometary material, on the
given phase, J is the gas flux (for details, see Prialnik et al. 1993), order of 104 dyn cm−2 (Klinger et al. 1989, Kochan et al. 1989,
and q is the sublimation rate. The porosity is given by Greenberg et al. 1995, Sirono and Greenberg 2000). For illus-
tration consider an element of porous solid material containing a

p =1− ρα,s / α − ρdust / dust , (5) fraction X of volatile ice, which absorbs heat continually. Even-
α tually, the ice will start sublimating and if the vapor is not allowed
to escape, the resulting gas pressure in the pores may rise up to
where α is the characteristic solid density. The energy conser- Rg XρT / pµ (the temperature will rise as well, controlled by
vation equation is the saturated vapor pressure, as in a pressure-cooker). This gas
  pressure exceeds the estimated material strength of cometary
∂    nuclei already for X ≈ 5 × 10−5 . Thus even a moderate build-
ρα u α + ∇ · F + u α Jα = λρa Hac + Q̇ − qα Hα , up of internal pressure is bound to open channels that would
∂t α α α
release the pressure, allowing the gas to escape. As a result of
(6) the easy escape of gas, the gas pressure in the porous interior
will be lower than the saturated pressure and therefore most of
subject to the following relations the heat released will be absorbed in sublimation.
 The effect of a few such scattered wide channels cannot, how-
µα ever, be accounted for by a 1-D code. We could, of course, adopt
qα = S(Pα (T ) − Pα ) (7)
2π Rg T a very large average channel width, but this would also imply an
 unrealistic surface-to-volume ratio, as well as an unrealistically
Q̇ = X 0,j Hj τj−1 exp(−t/τj ), (8)
small Knudsen number, and it would amplify heat advection.
j
Another approach is to assume that the effective permeability is
where u denotes specific heat, H is the latent heat, P is saturated sufficiently high to permit a quasi-steady-state approximation,
vapor pressure as given by the Clausius–Clapeyron equation, P where the first term on the left-hand side of the mass balance
is gas pressure, µ is molecular mass, Rg is the gas constant, S equation for the gas phases (Eq. (3)) is neglected, while at the
is the surface-to-volume ratio (depending on porosity and pore same time adopt a reasonable average pore-size. Thus Eqs. (3)–
size), Q̇ is the rate of radioactive energy release, including all the (4) are replaced by
isotopes of Table I, and τj is the decay time of the j th radioac-
tive isotope; X 0,j is its mass fraction, which for the long-lived ra- ∂ρα,s
∇ · Jα = qα , = −qα . (9)
dionuclides is obtained by the values listed in Table I, multiplied ∂t
by the dust mass fraction, while for 26 Al it is a free parameter.
Finally, Hj is the energy released per unit mass upon decay. The In this way we have to strictly solve only one time-dependent
heat flux is F = −ψ( p)K (T )∇T , where K (T ) is the thermal equation—although gas densities and production rates change as
conductivity of the solid material (as given by Klinger 1980 for the temperature distribution changes—supplemented by struc-
the ice component, while for dust K = 106 erg s−1 cm−1 K−1 ), ture (space-dependent) equations. This constitutes a huge com-
corrected by a factor ψ( p) = 1 − p 2/3 < 1 to take account of the putational advantage, since a long-term calculation including a
porosity. The boundary conditions are detailed account of gas flow through the porous medium, coupled
with heat transfer, would require a prohibitively large amount
F(0, t) = Jα (0, t) = 0 of computing time. Combining Eqs. (9) and integrating over
volume, we obtain
Pα (R, t) = 0
F(R, t) = σ T 4 (R, t) − (1 − A)L  /16πdH2 (t). − Ṁ α,s = Jα (R, t)4π R 2 , (10)
304 CHOI ET AL.

TABLE II sumed, at a heliocentric distance of 30 AU, although KBOs do


Composition Parameters (Given ρ = 0.7 g cm−3 ) not move in circular orbits; the exact orbit has no importance
in a long-term evolution calculation focused on the interior of
Mass Specific density Partial density Relative
the nucleus, and the distance chosen is smaller than the average
Component fraction g cm−3 g cm−3 volume
semi-major axis of ∼40 AU, since the most marked effect of
H2 O 0.40 0.917 0.280 0.305 irradiation is expected to occur near perihelion (i.e., closer than
CO 0.05 1.1 0.035 0.032 40 AU).
CO2 0.05 1.56 0.035 0.022 The initial 26 Al content is 5 × 10−8 , about 10 times lower
Dust 0.50 3.25 0.350 0.108
than the nominal value (see Section 2), allowing for a few Myr
Pores — — — 0.533
formation time. The models in Table III are grouped according to
different criteria: those in the first group differ from the baseline
model each in only one parameter. This allows an assessment
of the influence of each parameter. Models in the second group
which means that the total mass of gas ejected through the
share one parameter with the baseline model, while the other
comet’s surface per unit time is equal to the total amount of
two are determined by requirement (1), that is, equal power of
gas evaporated throughout the nucleus per unit time, for each
the two energy sources. The parameter combinations are self-
species. This approximation is valid for non-abundant species,
consistent: thus, model e has the same 26 Al content, but a smaller
for which the bulk density is low. It breaks down when the net
radius than a, implying a shorter formation time at the same
gas sublimation rate is negative, that is, when recondensation
distance, which is compensated by a larger heliocentric distance,
surpasses sublimation, but this is not the case for the conditions
where formation rates are lower. Model f has the same radius
considered in this paper. The calculations are terminated if vig-
as model a, but a larger distance and thus a lower 26 Al content;
orous water ice sublimation takes place.
model g has a larger radius at the same distance as model a,
It is worth noting that a different approximation with the same
compensated by a lower 26 Al content. The two models in the
computational advantage—reduction of the number of time-
third group are meant to provide a lower limit to the effect of
dependent equations—has been used in other studies (e.g., De
long-term evolution in the Kuiper Belt region.
Sanctis et al. 2001). It assumes that when both the ice and gas
In what follows we shall refrain from describing in detail the
phases are present, the gas density is equal to the saturated vapor
evolution of each model; instead, we shall focus on the broad
density, which is a function of temperature. Strictly, this would
view of their evolutionary pattern, looking, in particular, for
imply that no evaporation/condensation could take place. How-
properties and processes that appear to be shared by objects in
ever, as the temperature changes, the saturated density (pressure)
the Kuiper Belt. We shall compare models, in order to single out
changes with it, and this change can be translated into a rate of
the potential effects of individual parameters.
sublimation/condensation. This approximation is not valid for
The evolution of the central temperature Tc , the surface tem-
minor volatile components and fails close to the surface of the
perature Ts , and the maximal temperature Tmax attained within
nucleus, but it applies to water ice in the interior of the comet.
the nucleus is shown in Fig. 2 for models a–d. The abscissa in
The two simplifying approximations are thus complementary.
these (and the following) diagrams is time on a logarithmic scale,
The free parameters of our study are the cometary radius,
since changes are largely dictated by radioactive decay, which
the heliocentric distance (≥30 AU), and the initial content of
26 is exponential. Clearly, 26 Al plays a crucial role, raising the in-
Al. For all models we assume an initial temperature of 10 K,
ternal temperature to ∼180 K. (Calculations were terminated at
equal mass fractions of ices and dust, and a bulk density of
this point, when significant sublimation of water in the interior
0.7 g cm−3 (corresponding to a porosity of 0.5). The ice mix-
ture includes 10% (by mass) of each CO—representative of
extremely volatile species—and CO2 —representative of mod- TABLE III
erately volatile species—and the rest is amorphous water ice. Initial Parameters of Computed Models
Details of composition-related properties are summarized in
Model R (km) dH (AU) X 0 (26 Al)
Table II. We assume an albedo of 0.04 and an emissivity of
0.96, but the exact albedo value has little effect on the results, a 100 30 5(−8)
so long as it is significantly smaller than unity, which is the case b 100 30 0
for KBOs (Jewitt et al. 2001 and references therein). c 100 120 5(−8)
d 10 30 5(−8)
a 100 30 5(−8)
4. RESULTS OF EVOLUTIONARY CALCULATIONS
e 10 90 5(−8)
f 100 120 3(−9)
The eight models that were calculated, for different combi- g 500 30 1(−8)
nations of the three basic parameters, are listed in Table III. b 100 30 0
The baseline model is labeled a; it has a radius of 100 km—the h 10 30 0
typical value for radii of observed KBOs. A circular orbit is as-
THERMAL EVOLUTION IN THE KUIPER BELT 305

FIG. 2. Evolution of central, surface, and peak temperatures, Tc (dotted line), Ts (dashed line), and Tmax (solid line) for models a, b, c, d, respectively, as
marked. Note that the time scale is logarithmic, since changes are more pronounced during the early stages of evolution. The temporary halt in the rising trend
of the central temperature, first at ∼20 K and then again at ∼70 K is due to the sublimation of CO and CO2 , respectively, which absorbs the energy released by
radioactive decay. The plateau at ∼140 K is maintained by the crystallization of amorphous ice, which is exoergic.

was in conflict with the assumption of easy escape of gases—a the temperature peak moves with the crystallization front from
very large amount of vapor could not escape through very cold the center toward the surface. In all models there is a conspicu-
regions without considerable refreezing (see Section 3).) We ous temporary halt in the rising trend of the central temperature,
note the effect of radioactive 40 K causing a late rise in temper- first at ∼20 K and then again at ∼70 K. This is due to the subli-
ature in model b. We also note the effect of a small radius: in mation of CO and CO2 , respectively, which absorbs the energy
model d the maximal temperature is only ∼140 K. In all cases, released by radioactive decay.
the temperature rises rapidly at the surface, due to insolation, Similar conclusions are drawn from Fig. 3, which shows the
and more slowly in the interior. When the surface temperature evolution of Tc , Ts , and Tmax for models a and e–g. In addition,
reaches equilibrium with the solar radiation, it stops rising, and we note the effect of the initial amount of 26 Al: lowering the mass
the more slowly rising central temperature becomes the temper- fraction by an additional factor of ∼10 (to 3 × 10−9 —model f)
ature maximum. In models a and c (large radii and 26 Al con- reduces its effect significantly—the maximal internal tempera-
tent), the temperature peak shifts again outward after the 26 Al ture attained is only slightly above 50 K. Even for a large body,
decay. This is due to another energy source that has now been a reduction in the 26 Al content (model g) results in consider-
activated—the exoergic crystallization of amorphous ice. Thus ably lower temperatures, somewhat below 130 K. In this case,
306 CHOI ET AL.

FIG. 3. Same as Fig. 2 for models a, e, f, g.

however, the effect of the next potent radionuclide, 40 K, is ap- tained for a longer period of time in the slightly hotter object.
parent, causing the internal temperature to rise above 130 K Finally, when no 26 Al is present, the other radioactive isotopes
on a time scale of a few 108 years. We note that the evolu- have a limited effect on large bodies, as already mentioned, but
tionary course is little affected by heliocentric distance, except no effect at all on small bodies (∼10 km), as shown in Fig. 4 for
that at distances much larger than 30 AU (models c, e, f), the model h (cf. Consolmagno and Lewis 1978).
surface temperatures rise above equilibrium for a while, but re- The high internal temperatures that we have found to develop
main well below 50 K at all times. The effect of heliocentric in objects in the Kuiper Belt region are bound to lead to alter-
distance is more pronounced for a smaller body: this emerges ations in structure and composition. We proceed to illustrate the
from the comparison of model e with model d of the previous thermal and structural evolution of the models in color-coded
group—the peak temperature is lower in the more distant object 3-D plots. The ordinate represents the depth into the nucleus,
and it is maintained for a shorter time span. The reason is that again on a logarithmic scale, since usually gradients become
while both models attain sufficiently high internal temperatures steeper toward the surface. Figure 5, top and bottom, shows the
(130–140 K) for crystallization to occur, the small temperature evolution of temperature for models a–d and a ∪ e–g, respec-
difference between them results in a difference of a factor of 20 tively. We draw attention to model g, the 500-km object, which
in the exponential crystallization rate. Keeping in mind that crys- appears to maintain a high temperature in the interior, possi-
tallization is exoergic, this explains why the process—as well bly up to the present. Such a structure might have implications
as the relatively high temperature accompanying it—is main- for unusual activity patterns, such as are sometimes observed
THERMAL EVOLUTION IN THE KUIPER BELT 307

FIG. 4. Same as Fig. 2 for models b, h.

(Hainaut et al. 2000, Delsanti et al. 2001), if the relatively hot shown in Fig. 7, top and bottom, for models a–d and a ∪ e–g,
interior could be tapped (e.g., by a collision). respectively. The ordinate in these plots is the relative volume
Regarding the composition of the comet models, the changes (V (r )/V (R)) rather than the logarithm of depth, which means
are considerable, which should be expected in view of the high that the surface is now at the top. In addition, each panel has
temperatures attained. The initial content of volatiles besides its own gray scale, to emphasize individual details. When very
water corresponds to a bulk mass density of 0.07 g cm−3 (half CO little or no 26 Al is included and only the CO ice is lost, the re-
and half CO2 , see Table II). The change of this density through- sulting increase in porosity is small (about 7%, from 0.533 to
out the nucleus with time is shown in Fig. 6, top and bottom, 0.565—see Table II) and the porosity remains almost homoge-
for models a–d and a ∪ e–g, respectively. We draw attention to neous. When the CO2 ice is lost as well, regions of relatively
the obvious difference in the nature of the two energy sources: high and relatively low porosities are obtained due to evapora-
insolation, which is a surface source, generates a sublimation tion of water on the one hand and refreezing of CO2 on the other.
front of CO ice, which advances inward. This is illustrated by Thus the porosity ranges from as low a value as 0.2 to almost 0.8.
the slope of the boundary between the CO-rich and CO-depleted These changes should, however, be taken as indications of the
zones in all panels of Fig. 6. Radiogenic heating (of homoge- expected trends, since the assumption of very high permeabil-
neously distributed radioactives) is a volume source and affects ity is bound to affect these results far more than it has affected
most of the comet’s volume simultaneously; this is shown by the previous results. The range of possible porosities and bulk
the sublimation of CO2 , whose gradually decreasing density is densities that could arise from evolution of the initial structure
illustrated by the vertical, color-changing fronts in Fig. 6. The assumed (for all models) is summarized in Table IV. It is note-
CO ice is lost, sooner or later, in all models. The CO2 ice is worthy that large variations in porosity arise in a relatively thin
retained in a surface layer in all models, but the thickness of this region (in terms of radius), at depths ranging among models from
CO2 -rich layer varies considerably among them: all the CO2 is a few hundred meters to 1 km, which means that a mechanically
retained (everywhere) if no (or very little) 26 Al is included; only “weak” region forms at that depth below the surface. This may
the outer few hundred meters retain the CO2 ice in the base-
line model. Typically, the CO2 -rich layer is about 1 km thick,
almost regardless of the model’s radius, distance, and X 0 (26 Al), TABLE IV
> 10−8 ). In some cases, some of the CO2 gas
provided X 0 (26 Al ∼ Porosity and Density Ranges
released in the interior refreezes at the lower boundary of the Composition Porosity Density g cm−3
CO2 -rich outer layer, raising the CO2 density there by up to a
factor of 4. The region that retains the CO2 ice also retains the Dust + H2 O + CO2 + CO 0.533 0.70
water ice in amorphous form. Below it there is a ∼100-m thick Dust + H2 O + CO2 0.565 0.665
layer of amorphous ice, devoid of CO2 ice, and below that only Dust + H2 O 0.587 0.63
Dust 0.892 0.35
crystalline water ice and dust.
Dust + H2 O-filled pores 0 1.17
The loss of volatiles from the interior naturally affects the Dust + CO2 -filled pores 0 1.74
porosity. The evolution of porosity throughout the nucleus is
308 CHOI ET AL.

FIG. 5. Color map of the evolving temperature profile for models a, b, c, FIG. 6. Color map of the total mass density of CO and CO2 ices, as it
d—top panels, as marked—and models a, e, f, g—bottom panels, as marked. The changes with time throughout the nucleus, for the same models (as marked) and
change in color along a vertical line represents the radial temperature variation at on the same time and depth scales as those in Fig. 5. The initial CO + CO2 is
the corresponding time. The change in color along a horizontal line represents the 0.07 g cm−3 (see Table II); the density drops to 0.035 g cm−3 , when the CO
change of temperature with time at the corresponding (fixed) depth. The time ice has completely evaporated and only the CO2 ice is left, and drops to zero,
and depth scales are logarithmic, since changes are more pronounced during when both these volatile species have evaporated. Density values higher than
the early stages of evolution, and sharper, generally, toward the surface of the 0.07 g cm−3 result from refreezing of CO2 gas, as it flows outward through cold
nucleus. layers.
THERMAL EVOLUTION IN THE KUIPER BELT 309

different assumptions regarding physical characteristics, but re-


taining the same initial porosity. The conclusions of these tests
may be summarized as follows:

(a) When gases trapped in amorphous ice are considered, in


addition to those included in frozen form, the only long-term
change is that more volatiles can freeze, eventually, in the outer
layers of the nucleus, reducing the porosity. However, this effect
depends entirely on the initial amounts of trapped and frozen
volatiles and it would be impossible to deduce the origins of
the ice in the outer layers of the nucleus based on the final
composition there.
(b) We have run models for which we solved the full set of
time dependent equations, adopting a very high permeability
(large average pore size)—an option mentioned in Section 3—
for comparison. The grid spacing in the outer part of such models
is far coarser, since the mass-shell thickness is limited from be-
low by the pore size. Therefore the surface temperature is less
accurate. As expected, advection surpasses conduction, some-
times by a large factor, in this case. Nevertheless, we do not find
a marked difference in the thermal evolution of these models
as compared to the former models. Our calculations show that
among the different terms of the heat equation, conduction and
advection are smaller than the other terms during the decay of
26
Al, meaning that the energy released is mainly absorbed lo-
cally, in evaporation. This strengthens our conclusions regarding
the internal structure of KBOs.
(c) For the test models we also find that larger amounts of gas
flowing outward from the interior freeze in the cold outer regions,
reducing the porosity there to very low values. For example, for a
500-km model (similar to model g) a porosity of only about 0.06
may be reached after 1 Myr of evolution, in an outer layer about
1 km thick. This is because the internal gas densities are higher,
meaning that a gas molecule dwells in the nucleus interior for
a longer period of time, which increases the probability of its
refreezing.
(d) Increasing the baseline heliocentric distance from 30 to
40 AU has no significant effect. Although the surface tempera-
ture is lower, we find that the volatile CO is still completely lost.
We should mention here that the heliocentric distance adopted
FIG. 7. Gray-scale map of the changing porosity for models a, b, c, d—top is, in fact, only a measure of the solar radiation. The true pa-
panels, as marked—and models a, e, f, g—bottom panels, as marked—where the rameter of these models is L  /dH2 . Since the solar luminosity
ordinate is relative volume, rather than depth, providing a different perspective of was weaker in the early stages of the Solar System, whereas we
the models’ structures. It is meant to indicate the mass distribution (since porosity
have assumed here the present value of L  , the true distances
is a volume-related parameter). Individual scale bars are used in this case.
corresponding to our nominal distances may, in fact, be smaller
by as much as 15%. Hence the fact that a 25% change in distance
be significant for the effect of collisions, which are relatively has no significant effect is encouraging.
frequent in the Kuiper Belt, bearing on the sizes of broken-off
fragments. A detailed comparison with other similar studies is very dif-
So far, we have performed a systematic parameter study, in- ficult, if not impossible, to perform for two reasons: first, the
volving only three parameters. In fact, however, the number of approximations and basic assumptions employed differ among
uncertain parameters is larger and it would be impossible to ac- models, and second, the free parameter values adopted are dif-
count for all of them in a methodical manner. Therefore, in order ferent. Nevertheless, we find that, broadly, the results of different
to test the robustness of our results regarding the change of in- calculations agree and complement each other. As mentioned in
ternal structure, we have computed additional test models, with Section 1, a calculation similar to the present calculation was
310 CHOI ET AL.

recently performed by De Sanctis et al. (2001), for two particu- TABLE V


26
lar KBO orbits and radii and for a limited evolution time. As a Al Required for Evaporation of Volatiles
result, the effect of radioactive heating was less apparent in that
Molecule Tsubl (K) H (erg g−1 ) X 0 (26 Al)
study. Depletion of volatiles, differentiation of the initially ho-
mogeneous composition, and refreezing are the main results of H2 O 120 2.8(10) 1.0(−7)
both these calculations. Our results are also in agreement with CO 20 2.3(9) 8.2(−9)
those of the extensive parameter study performed by Prialnik CO2 70 5.9(9) 2.4(−8)
and Podolak (1995), although that study was devoted mainly NH3 83 1.7(10) 6.2(−8)
CH4 30 6.2(9) 2.2(−8)
to the process of crystallization and release of gas trapped in
H2 CO 73 8.2(9) 3.1(−8)
amorphous ice, whereas here volatile species are included as ice N2 21 2.5(9) 8.9(−9)
mixtures.

5. CONCLUSIONS 3. KBOs may have partially lost less volatile ices as well. The
temperatures at which sublimation of different volatiles sets in
We have carried out long-term evolutionary calculations for (Tsubl ), the corresponding latent heat, and the mass fraction of
models simulating potential KBOs. A composition of porous ice 26
Al that would be required in order to raise the temperature of
and dust was assumed, where the ice was a mixture of H2 O—the cometary material from initially 10 K to Tsubl and sublimate var-
most abundant species—and CO and CO2 in smaller amounts. ious volatiles, assuming they constitute 10% of the ice, are listed
The main purpose of the study was to determine whether these in Table V. These are, of course, lower limits to the necessary
more volatile species could survive radioactive heating during amounts of 26 Al, since some of the heat is conducted outward
the early stages of evolution and whether a presumably homo- and “wasted.” Thus, CO, as well as N2 and possibly methane, are
geneous initial structure and composition could be preserved. expected to be lost entirely, but the moderately volatile species,
To account for large vents or channels that should allow gases such as CO2 , H2 CO, and NH3 , should be partially retained. As
to escape—and also in order to facilitate the computations—we for water, only a relatively small fraction could be evaporated.
have adopted a quasi-steady-state approximation for the gaseous 4. As a result, the internal structure of KBOs is most prob-
phases. Spherical symmetry was also assumed, but this is per- ably not uniform; rather, density, porosity, H2 O ice phase, and
fectly justified in long-term evolution calculations, especially strength, vary with depth. Generally, the porosity should increase
when the major heat source is homogeneously distributed over with depth (since compaction is unlikely to occur in small bod-
the mass. Based on these assumptions and on the set of param- ies), but not necessarily monotonically. In particular, we have
eters adopted, and allowing for the approximations employed, found that weak regions—by which we mean regions of sharp
our main conclusions are as follows: density variations—may form at depths ranging among models
from several hundred meters to 1 km. Since breaking occurs
1. The internal temperature profile of KBOs may have been preferentially at, or along, a weak zone, it is reasonable to de-
substantially affected by both short- and long-lived radionu- duce that fragments breaking off KBOs upon collision are likely
clides, with accompanying changes in composition and struc- to have sizes on the order of 1 km. Now, collisions in the Kuiper
ture. The evolution of the temperature profile and the structural Belt are believed to be responsible for the Jupiter-family comets
modifications are a strong function of the accretion times of and thus, indulging in speculation, we may have found a reason
KBOs, their sizes, the dust-to-ice mass fraction, and, to some for the relatively small sizes of these comets (see also Farinella
extent, the heliocentric distances. and Davis 1996, Davis and Farinella 2001).
2. KBOs may have lost entirely all volatiles that sublimate 5. Not only is the structure not uniform, but also, similarly,
below ∼40–50 K, which were initially included as ices. This the internal composition of KBOs is most probably not homo-
can occur even in the absence of 26 Al. However, in this case the geneous, but stratified (see also De Sanctis et al. 2001), with the
conclusion is valid only on the assumption that the entire surface outer layers being less altered by evolution. We have found that
area is—on average—equally heated. Depending on the angle of regions enriched in volatile species, as compared with the initial
the rotation axis of the object, it may well be that some areas are abundances assumed, arise due to gas migration and refreezing.
always concealed from the Sun and stay at very low temperatures This effect should be more pronounced, if the assumption of
at all times. But it seems unreasonable to assume that rotation of very high permeability (in different approximations) is relaxed.
a small (and probably uneven) body would remain undisturbed When such regions are exposed in comets originating in the
for billions of years. Thus, it is highly unlikely that KBOs still Kuiper Belt, after the overlying material has been eroded, vigor-
retain extremely volatile species in icy form. This leads to the ous evaporation of these volatile species may result in outbursts.
further conclusion that if comets originating in the Kuiper Belt
emit such volatiles (e.g., CO) these must have been trapped and Our results should be taken to indicate evolutionary trends and
retained in the amorphous ice, meaning that the ice of KBOs our conclusions have been formulated as qualitative. An uneven
must have been initially amorphous. shape, rotation, orbital migration, and collisions are only a few
THERMAL EVOLUTION IN THE KUIPER BELT 311

among the factors that may affect the structure of KBOs, but are Greenberg, J. M., H. Mizutani, and T. Yamamoto 1995. A new derivation of the
difficult—if not impossible—to be accounted for in a systematic tensile strength of cometary nuclei: Application to Comet Shoemaker–Levy 9.
manner. Nevertheless, we may state with a reasonable degree Astron. Astrophys. 295, L35–L38.
of confidence that if KBOs did experience radioactive heating, Hainaut, O. R., C. E. Delahode, H. Boehnardt, E. Dotto, M. A. Barucci,
K. J. Meech, J. M. Bauer, R. M. West, and A. Doressoundiram 2000. Physical
their structure and composition were altered mainly to the extent properties of TNO 1996 TO66 —Lightcurves and possible cometary activity.
of considerable loss of volatiles and significant departure from Astron. Astrophys. 356, 1076–1088.
internal homogeneity. Haruyama, J., T. Yamamoto, H. Mizutani, and J. M. Greenberg 1993. Thermal
history of comets during residence in the Oort cloud: Effect of radiogenic
ACKNOWLEDGMENTS heating in combination with the very low thermal conductivity of amorphous
ice. J. Geophys. Res. 98, 15,079–15,088.
We thank Paul Weissman for a very careful reading of the original manuscript Jewitt, D., J. Luu, and C. Trujillo 1998. Large Kuiper Belt objects: The Mauna
and for many valuable comments an suggestions. R.M. acknowledges the support Kea 8K CCD survey. Astron. J. 115, 2125–2135.
of the Minerva Foundation. Jewitt, D., H. Aussel, and A. Evans 2001. The size and albedo of the Kuiper
Belt Object (2000) Varuna. Nature 411, 446–447.
REFERENCES Kenyon, S. J., and J. X. Luu 1998. Accretion in the early Kuiper Belt. I. Coag-
ulation and velocity evolution Astron. J. 115, 2136–2160.
Akridge, G., P. H. Benoit, and D. W. G. Sears 1998. Regolith and megaregolith Klinger, J. 1980. Influence of a phase transition of ice on the heat and mass
formation of H-chondrites: Thermal constraints on the parent body. Icarus balance of comets. Science 209, 271–272.
132, 185–195. Klinger, J., S. Espinasse, and B. Schmidt 1989. Some considerations on cohesive
Brown, R. H., and C. C. Koresko 1998. Detection of water ice on the Centaur forces in sun-grazing comets. Europ. Space Agency Spec. Publ. 12, 197–200.
1997 CU26. Astrophys. J. 505, L65–L67. Kochan, H., B. Feuerbacker, F. Joo, J. Klinger, W. Sebolt, A. Bischoff, H. Duren,
Brown, R. H., D. P. Cruikshank, Y. Pendleton, and G. J. Veeder 1998. Iden- D. Stuffler, T. Spohn, H. Fechtig, E. Gruen, H. Kohl, D. Krankowsky, K.
tification of water ice on the Centaur 1997 CU26. Science 280, 1430– Roessler, K. Thiel, G. Schwehm, and U. Weishaupt 1989. Comet simulation
1432. experiments in the DFVLR space simulators. Adv. Space Res. 9, 113–121.
Brown, R. H., D. P. Cruikshank, and Y. Pendleton 1999. Water ice on Kuiper Lee, T., D. Papanastassiou, and G. J. Wasserburg 1976. Demonstration of 26 Mg
Belt Object 1996 TO66 . Astrophys. J. 519, L101–L104. excess in Allende and evidence for 26 Al. Geophys. Res. Lett. 3, 109–112.
Chiang, E. I., and M. E. Brown 1999. Keck pencil-beam survey for faint Kuiper Luu, J. X., D. C. Jewitt, and C. Trujillo 2000. Water ice in 2060 Chiron and its
Belt objects. Astron. J. 118, 1411–1422. implications for Centaurs and Kuiper Belt objects. Astrophys. J. 531, L151–
Clayton, D. D., and M. D. Leising 1987. 26 Al in the interstellar medium. Phys. L154.
Rep. 144, 1–50. MacDonald, G. J. 1959. Calculations on the thermal history of the Earth. J. Geo-
Consolmagno, G. J., and J. S. Lewis 1978. The evolution of icy satellite interiors phys. Res. 64, 1967–2000.
and surfaces. Icarus 34, 280–293. Mahoney, W. A., J. C. Ling, Wm. A. Wheaton, and A. S. Jacobson 1984. Heao
Davis, D. R., and P. Farinella 1997. Collisional evolution of Edgeworth–Kuiper 3 discovery of 26 Al in the interstellar medium. Astrophys. J. 286, 578–585.
Belt Objects. Icarus 125, 50–60. Mekler, Y., D. Prialnik, and M. Podolak 1990. Evaporation from a porous comet
Davis, D. R., and P. Farinella 2001. Collisional effects in the Edgeworth– nucleus. Astrophys. J. 356, 682–686.
Kuiper Belt. In Collisional Processes in the Solar System (M. Ya. Marov and Miyamoto, M., N. Fujii, and H. Takeda 1981. Ordinary chondrite parent body:
H. Rickman, Eds.), pp. 277–285. Kluwer, Dordrecht. An internal heating model. Proc. Lunar Planet. Sci. Conf. 12B, 1145–1152.
Delsanti, A. C., H. Boehnardt, L. Barrera, K. J. Meech, T. Sekiguchi, and O. R. Orosei, R., A. Coradini, M. C. De Sanctis, and C. Federico 2001. Collision-
Hainaut 2001. BVRI photometry of 27 Kuiper Belt objects with ESO/very induced thermal evolution of a comet nucleus in the Edgeworth–Kuiper Belt.
large telescope. Astron. Astrophys. 380, 347–358. Adv. Space Res. 28, 1563–1569.
De Sanctis, M. C., M. T. Capria, and A. Coradini 2001. Thermal evolution and Podolak, M., and D. Prialnik 1997. 26 Al and liquid water environments in comet.
differentiation of Edgeworth–Kuiper Belt objects. Astron. J. 121, 2792–2799. In Comets and the Origin of Life (P. Thomas, C. Chyba, and C. McKay, Eds.),
Duncan, M., T. Quinn, and S. Tremaine 1988. The origin of short-period comets. pp. 259–272. Springer-Verlag, New York.
Astrophys. J. 328, L69–L73. Pollack, J. B., O. Hubickyj, P. Bodenheimer, J. J. Lissauer, M. Podolak, and
Ellsworth, K., and G. Schubert 1983. Saturn’s icy satellites: Thermal and struc- Y. Greenzweig 1996. Formation of the giant planets by concurrent accretion
tural models. Icarus 54, 490–510. of solids and gas. Icarus 124, 62–85.
Farinella, P., and D. R. Davis 1996. Short-period comets: Primordial bodies or Prialnik, D. 1992. Crystallization, sublimation, and gas release in the interior of
collisional fragments. Science 273, 938–941. a porous comet nucleus. Astrophys. J. 388, 196–202.
Farinella, P., D. R. Davis, and S. A. Stern 2000. Formation and collisional Prialnik, D., and A. Bar-Nun 1990. Heating and melting of small icy satellites
evolution of the Edgeworth–Kuiper Belt. In Protostars and Planets IV by the decay of 26 Al. Astrophys. J. 355, 281–286.
(V. Mannings, A. P. Boss, and S. S. Russell, Eds.), pp. 1255–1282. Univ. Prialnik, D., and M. Podolak 1995. Radioactive heating of porous cometary
of Arizona Press, Tucson. nuclei. Icarus 117, 420–430.
Ghosh, A., and H. Y. McSween 1998. A thermal model for the differentiation Prialnik, D., and M. Podolak 1999. Changes in the structure of comet nuclei due
of Asteroid 4 Vesta, based on radiogenic heating. Icarus 134, 187–206. to radioactive heating. Space Sci. Rev. 90, 169–178.
Gladman, B., and M. Duncan 1990. On the fates of minor bodies in the outer Prialnik, D., A. Bar-Nun, and M. Podolak 1987. Radiogenic heating of comets
Solar System. Astron. J. 100, 1680–1693. by 26 Al and implications for their time of formation. Astrophys. J. 319, 993–
Gladman, B., J. J. Kavelaars, P. D. Nicholson, T. J. Loredo, and J. A. Burns 1002.
1998. Pencil-beam surveys for faint trans-neptunian objects. Astron. J. 116, Prialnik, D., U. Egozi, A. Bar-Nun, M. Podolak, and Y. Greenzweig 1993. On
2042–2054. pore size and fracture in gas-laden comet nuclei. Icarus 106, 499–507.
312 CHOI ET AL.

Schmitt, B., S. Espinasse, R. J. A. Grim, J. M. Greenberg, and J. Klinger 1989. Urey, H. C. 1955. The cosmic abundances of potassium, uranium and thorium
Laboratory studies of cometary ice analogues. Europ. Space Agency Spec. and the heat balances of the Earth, the Moon and Mars. Proc. Nat. Acad. Sci.
Publ. 302, 65–69. 41, 127–144.
Schubert, G., D. J. Stevenson, and K. Ellsworth 1981. Internal structures of the Wallis, M. K. 1980. Radiogenic heating of primordial comet interiors. Nature
Galilean satellites. Icarus 47, 46–59. 284, 431–433.
Share, G. H., R. L. Kinzer, J. D. Kurfess, D. J. Forrest, E. L. Chupp, and E. Rieger Weidenschilling, S. J. 1988. Formation processes and time scales for me-
1985. Detection of galactic 26 Al gamma radiation by the SMM spectrometer. teorite parent bodies. In Meteorites and the Early Solar System (J. F.
Astrophys. J. 292, L61–L65. Kerridge and S. W. Metthews, Eds.), pp. 348–371. Univ. of Arizona Press,
Sheppard, S. S., D. C. Jewitt, C. Trujillo, M. J. I. Brown, and M. C. B. Ashley Tucson.
2000. A wide-field CCD survey for centaurs and Kuiper Belt objects. Astron. Weidenschilling, S. J. 1997. The origin of comets in the solar nebula: A unified
J. 120, 2687–2694. model. Icarus 127, 290–306.
Sirono, S., and J. M. Greenberg 2000. Do cometesimal collisions lead to bound Wetherill, G. W., and G. R. Stewart 1989. Accumulation of a swarm of small
rubble piles or to aggregates held together by gravity? Icarus 145, 230–238. planetesimals. Icarus 77, 330–357.
Srinivasan, G., J. N. Goswami, and N. Bhandari 1999. 26 Al in Eucrite Piplia Whipple, F. L., and R. P. Stefanic 1966. On the physics and splitting of cometary
Kalan: Plausible heat source and formation chronology. Science 284, 1348– nuclei. Mem. R. Soc. Liege (Ser. 5) 12, 33–52.
1350. Yabushita, S. 1993. Thermal evolution of cometary nuclei by radioactive heating
Trujillo, C., J. X. Luu, A. S. Bosh, and J. L. Elliot 2001. Large bodies in the and possible formation of organic chemicals. Mon. Not. R. Astron. Soc. 260,
Kuiper Belt. Astron. J. 122, 2740–2748. 819–825.

You might also like