Where Are The Extrasolar Mercuries?

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Draft version August 14, 2020

Typeset using LATEX twocolumn style in AASTeX63

Where are the Extrasolar Mercuries?


Alexandra E. Doyle, Beth Klein,2 Hilke E. Schlichting,1, 2, 3 and Edward D. Young1
1

1 Earth, Planetary, and Space Sciences


University of California, Los Angeles
Los Angeles, CA 90095, USA
arXiv:2008.05992v1 [astro-ph.EP] 13 Aug 2020

2 Physics and Astronomy

University of California, Los Angeles


Los Angeles, CA 90095, USA
3 Earth, Atmospheric and Planetary Sciences

Massachusetts Institute of Technology


Cambridge, MA 02139, USA

(Accepted ApJ, August 8, 2020)

ABSTRACT
We utilize observations of 16 white dwarf stars to calculate and analyze the oxidation states of the
parent bodies accreting onto the stars. Oxygen fugacity, a measure of overall oxidation state for rocks,
is as important as pressure and temperature in determining the structure of a planet. We find that
most of the extrasolar rocky bodies formed under oxidizing conditions, but ∼1/4 of the polluted white
dwarfs have compositions consistent with more reduced parent bodies. The difficulty in constraining the
oxidation states of relatively reduced bodies is discussed and a model for the time-dependent evolution
of the apparent oxygen fugacity for a hypothetical reduced body engulfed by a WD is investigated.
Differences in diffusive fluxes of various elements through the WD envelope yield spurious inferred bulk
elemental compositions and oxidation states of the accreting parent bodies under certain conditions.
The worst case for biasing against detection of reduced bodies occurs for high effective temperatures.
For moderate and low effective temperatures, evidence for relatively reduced parent bodies is preserved
under most circumstances for at least several characteristic lifetimes of the debris disk.

Keywords: white dwarfs: exoplanets

1. INTRODUCTION polluted white dwarf (WD) stars. White dwarfs are the
The characterization of rocky exoplanets is a grow- last observable stage of stellar evolution for stars where
ing area of research. Still, relatively little is known M∗ < 8M . Because of the extraordinary gravity as-
about the compositions of extrasolar rocky bodies, in sociated with these electron degenerate stars, elements
part because it is the atmospheres of exoplanets that are heavier than helium sink rapidly below their surfaces.
amenable to direct observations, and compositions de- Yet, spectroscopic studies show that 25 - 50% of WDs
duced from mass-radius relationships are highly degen- are ‘polluted’, exhibiting elements heavier than helium
erate and have limited precision (e.g. Dorn et al. 2015). (Zuckerman et al. 2003, 2010; Koester et al. 2014). The
A powerful technique for acquiring detailed geochemical source of these heavy elements in WDs is exogenous, the
information about rocky exoplanets is the analysis of result of accretion of rocky debris from parent bodies
that previously orbited the WDs (Jura & Young 2014;
Veras 2016; Farihi 2016; Zuckerman & Young 2017). Di-
Corresponding author: Alexandra E. Doyle rect evidence for the pollution being of planetary origin
a.doyle@ucla.edu exists in the form of infrared (IR) excesses in the spec-
Corresponding author: Edward D. Young tral energy distributions (SEDs) of the white dwarfs
eyoung@epss.ucla.edu (Zuckerman & Becklin 1987; Becklin et al. 2005; Jura
et al. 2007). The IR excesses emanate from circumstel-
2 Doyle et al.
100%
Al
90% Ca
Si
80% Mg
Fe
70%
O
60%
Mass

50%

40%

30%

20%

10%

0%
y e 3 1 5 6 75
r h e
la lley RB BSE art cur drit drit 855 520 40 83 11
5 0 17 12 45 -38 23 162 56 12 25 1-
So Ha MO E e r n
o n + 0 6 + D + 1 6+ 5+0 9+0 on 3 29 +80 0- 2+ 7+ 46- 24 1+1
et lk M ho h 43 53 G 8 22 4 2 T G 53 35 23 20
04
G 55
Bu C IC 0 1 73 1 11 19 1
22 D D 1 D 2 D 1
om tite C J1 WD J0 WD D D S W W D
C ta S S W W H W W W
s S S
En SD SD

Figure 1. Mass concentrations of the 6 major rock-forming elements in the 16 WDs used in this study normalized to 100%.
Generally, rocky bodies accreting onto these stars resemble rocks in our own Solar System and range from basaltic, crust-like
to bulk-asteroid or planet-like compositions. The excess oxygen seen in WD 1232+563 is discussed in Section 3.2. Note WD
1145+017 is updated from Doyle et al. (2019) using new abundance estimates from Fortin-Archambault et al. (2020). Solar
System objects are shown for comparison on the left.

lar dust disks that thermally reprocess the light from (Figure 1). While many of the observations have so far
the star (Jura 2003). It is estimated that ∼4% of white been consistent with dry, rocky parent bodies, there are
dwarfs have detectable IR excess attributed to circum- a few important exceptions. One WD exhibits a pattern
stellar debris disks (Kilic et al. 2008; Farihi 2009; Barber of abundances with unusually high O, C, and N that is
et al. 2012). An even smaller number of white dwarfs best explained as an extrasolar Kuiper-Belt-analog (Xu
exhibit disks of gaseous material in Keplerian rotation et al. 2017), and more recently, a white dwarf that is
(e.g. Gänsicke et al. 2006) that has been photoionised accreting material consistent with a giant planet in a
by the white dwarf (Melis et al. 2010). Other direct ev- close orbit around the star has been observed (Gnsicke
idence for the pollution being of planetary or asteroidal et al. 2019). Also, a handful of WDs exhibit evidence for
origin comes from detection of rapidly evolving tran- accretion of water-rich parent bodies (Jura 2008; Jura
sits, although to date only a handful of white dwarfs are et al. 2009; Jura & Xu 2010; Farihi et al. 2013; Raddi
observed to have a transiting body (Vanderburg et al. et al. 2015). This ability to classify the geochemistry
2015; Manser et al. 2019; Vanderbosch et al. 2019). of rocky extrasolar bodies can be taken to even greater
Spectroscopic studies of polluted WDs allow us to an- specificity by calculating the oxidation states of the par-
alyze the elemental constituents of extrasolar rocky bod- ent bodies (Doyle et al. 2019).
ies and evaluate the geochemical and geophysical nature Oxygen fugacity (fO2 ) is the effective partial pressure
of exoplanets in general. Indeed, a number of authors of O2 in a system and a commonly used measure of over-
have applied these principles to the dozen or so cur- all oxidation state for rocks. Oxygen fugacity is as im-
rently well-studied polluted WDs where the four ma- portant as pressure and temperature in determining the
jor rock-forming elements are detected and quantified characteristics of a rocky planet. It determines the min-
(Klein et al. 2010). It is possible that WDs are accret- erals comprising the rocky mantle of a planet and the
ing multiple small planetary bodies (e.g. Jura 2003). In abundance of light elements in its metallic core. These
these cases, mixing of material would minimize com- in turn are important factors in determining whether a
positional extrema. In this work, we assume pollution body will have a magnetic field and how much water
occurs from the disruption of single bodies. Generally can be stored in the interior (Pearson et al. 2014; Frost
speaking, the compositions of the bodies polluting WDs et al. 2001; Frost & McCammon 2008; Wood et al. 2006;
resemble those of rocky bodies in our own solar system Buffett 2000; Elkins-Tanton & Seager 2008). The miner-
Where are the Extrasolar Mercuries? 3

alogy and water content of the mantle will influence the enstatite chondrites formed at oxygen fugacities consis-
likelihood for plate tectonics and the nature of volcan- tent with those of the protoplanetary disk (with ∆IW
ism that contributes to the atmosphere (e.g. Kolzenburg ∼ −3 to −7; e.g. Cartier & Wood 2019), the majority
et al. 2018; Schaefer & Fegley 2017). of rocky bodies in the Solar System require an oxidiz-
Oxygen fugacity can always be defined thermodynam- ing mechanism that increases fO2 by approximately 5
ically to characterize the oxidation state of a rock by orders of magnitude (e.g. Monteux et al. 2018).
evaluating a thermodynamic expression for a reaction In this work we present the intrinsic oxygen fugacities
between rock components and O2 . Oxygen fugacities of for 16 rocky parent bodies that are accreting, or did ac-
planetary materials are commonly expressed relative to crete, onto white dwarfs. We evaluate the robustness of
that for the Iron-Wüstite equilibrium reaction between the calculated oxygen fugacities and determine the con-
pure metallic iron (Fe) and pure wüstite (FeO). This ditions under which reduced bodies could be observed.
so-called Iron-Wüstite (IW) reference reaction is The calculation method for ∆IW is described in Section
2. The oxygen fugacity results and a description of the
1 modeling efforts to evaluate the robustness of the ∆IW
Fe + O2
FeO. (1)
2 values is described in Section 3. In Section 4 we discuss
Relating the oxygen fugacity of a rock to a reference our results and investigate the likelihood of observing a
reaction like Equation 1 allows one to refer to oxygen reduced body like Mercury using our method. Section
fugacity by difference, ∆IW, such that 5 provides a brief summary of the conclusions for this
study.
∆IW = log (fO2 )rock − log (fO2 )IW . (2)
2. METHODS
Presenting oxygen fugacity in this way allows for the
comparison of values independent of temperature and 2.1. Selection of White Dwarfs
pressure, because the standard-state Gibb’s free energy In Doyle et al. (2019) we used stringent requirements
of the reaction that contains the pressure and tempera- for the WDs chosen for study. We required that each
ture dependence cancels in Equation 2. Only the equilib- star had quantifiable abundances of O, Fe, Mg, Si, Ca
rium constant composed of the activities of FeO and Fe and Al in their atmospheres and that each star had a
is required to specify the relative fO2 . Comparison be- debris disk as confirmed by the presence of an IR excess
tween different rocks is facilitated by setting the uncer- in the SED. In this work we loosen these constraints
tain activity coefficients to unity, as is common in plane- to increase the number of polluted WDs available for
tary science. The activity coefficients are effectively the estimating oxygen fugacities of accreted parent bodies.
same for all of the rocks and metals considered here, so Here we use stars with quantifiable abundances of O, Fe,
applying plausible activity coefficients would shift all of Mg, Si and Ca to calculate fO2 . We relax the require-
the ∆IW values shown here systematically by at most ment for data for the abundance of Al. Where only an
∼ one log unit. Therefore, the initial oxidation state of upper limit for Al is reported, or where no value is given
a rocky body with at least some iron metal at the time at all, we use the Al/Ca ratio for CI chondrites to scale
of its formation is recorded by the concentration of FeO the elemental abundance of Al to the Ca abundance in
in the rock (xrock
FeO ) and the concentration of Fe in the the WD (Lodders 2003). We scale to Ca because Al
metal (xmetal
Fe ), yielding and Ca have the same volatility in rock. In each of
 rock  these instances, we assign the element a measurement
xFeO
∆IW = 2log . (3) uncertainty of 0.301 dex, the value obtained by scaling
xmetal
Fe the measurement by a factor of 2. Aluminum is not as
Because Equation 3 refers to metal-silicate differenti- abundant in rock as the other major elements, and vary-
ation, it records the oxygen fugacity at the time the ing the assumed abundances by a factor of 2 alters ∆IW
planet or planetesimal was forming (differentiating), and values by 0.1 dex or less.
is what we refer to as the “intrinsic” oxygen fugacity of We now also include stars that do not have con-
a body. firmed debris disks. In the current study, 63% of the
In the Solar System, most rocky bodies formed under WDs (100% of the Hydrogen-dominated WDs and 50%
oxidizing conditions consistent with ∆IW ≥ −2 (e.g. of the Helium-dominated WDs) have confirmed IR ex-
Grossman et al. 2012). However, the oxygen fugacity of cesses indicative of debris disks. Two additional Helium-
a protoplanetary disk, a hydrogen-rich gas of solar com- dominated WDs have unconfirmed IR excesses from
position, is reduced with ∆IW ∼ −6 to −7 (Krot et al. WISE data, but WISE has been known to give false
2000). This dichotomy means that while Mercury and positives (Dennihy et al. 2020). Compared to the total
4 Doyle et al.

fraction of polluted WDs with debris disks, (∼4%; Bar- We propagate measurement uncertainties for the pol-
ber et al. 2012), our sample of WDs preferentially has luted WDs using a Monte Carlo bootstrap method with
material in debris disks that has yet to be accreted. replacement for each elemental abundance. The abun-
dance data are generally reported as log(Z/X) (Doyle
2.2. fO2 Calculation
et al. 2019) where Z is the element of interest (O, Fe,
In order to calculate oxygen fugacities, we follow the Mg, Si, Ca or Al) and X is the dominant atmospheric
methods described by Doyle et al. (2019). From the element for the WD (H or He). We draw single values
element abundance ratios, we assign oxygen to Mg, Si, at random for each element from the lognormal distri-
Ca, and Al in the necessary proportions to obtain the butions for Z/X as suggested by the symmetrical uncer-
relative abundances of the charge-balanced rock-forming tainties in the log ratio data. A single random draw for
oxide components MgO, SiO2 , CaO, and Al2 O3 . The each of the elements represents a single instance of the
remaining excess oxygen (Oxs ), is assigned to Fe to make WD data. We calculate a ∆IW value from this draw,
FeO until either O or Fe is exhausted. The excess oxygen and repeat the calculation 100,000 times to obtain a fre-
available to make FeO is obtained using quency distribution of ∆IW values. We report median
values for ∆IW and calculate asymmetrical uncertain-
Oxs OSiO2 OMgO OAl2 O3 OCaO ties based on an extension of the interquartile range to
=1− − − − , (4) encompass 67% of the distribution (Doyle et al. 2019).
OTotal OTotal OTotal OTotal OTotal
We check our methods by creating fictive polluted
where Oi is the amount of oxygen needed to form the
WDs in which element ratios for various Solar System
metal oxide, i, and OTotal is the total abundance of oxy-
rocky bodies are assigned abundances relative to He or H
gen in the system. Other studies have used similar meth-
consistent with typical observations (in this case, using
ods for budgeting oxygen (Klein et al. 2010, 2011; Farihi
GD 40 as the hypothetical WD target). The rocky, Solar
et al. 2011, 2013, 2016; Raddi et al. 2015). Once the rel-
System objects used in this study include bulk silicate
ative abundances of the oxides are obtained, they are
Earth (McDonough 2003), bulk silicate Mars (Taylor
normalized to 1, yielding mole fractions, and permitting
2013), bulk silicate Mercury (Nittler et al. 2019), Vesta
application of Equation 3. In principle, if insufficient
(Steenstra et al. 2016), chondrites (CI (Lodders 2003),
oxygen exists to pair with Fe to make FeO, the Fe that
CM, CV, H, L, LL and EH-EL silicate (Newsom 1995)).
remains should have been present as metal in the ac-
We assign uncertainties to these fictive log(Z/X) values
creted parent body. We emphasize that oxygen fugacity
based on those for GD 40 and propagate measurement
is recorded by the mole fraction of FeO which depends
uncertainties in the same way as described above. We
on all of the oxides (FeO, SiO2 , MgO, Al2 O3 , CaO), and
define the value for Oxs /Fe for each body to that of GD
not simply the FeO/Fe ratio for the body. It is possi-
40. This value is generally representative of those for
ble for metal and water to have coexisted in the parent
most of the WDs in this study (Table 1). These calcula-
body if it were undifferentiated, meaning that oxygen
tions provide us with a test of the accuracy of the various
which is attributed to FeO in this calculation may have
computational methods described here. The calculated
existed as H2 O in the parent body. However, during
oxygen fugacities for the fictive polluted WDs accreting
the differentiation of a rocky body, the oxygen from ices
Solar System objects are comparable to the canonical
will oxidize metallic Fe to form FeO. We are assuming
oxygen fugacities for these objects (e.g. Cartier & Wood
here that the bodies we are observing in the WDs were
2019).
either differentiated themselves, or they are the build-
ing blocks of differentiated bodies (chondrite meteorites
would be the appropriate analog). Where a parent body
was composed in part of Fe metal and H2 O, our calcula- 2.3. Constraining Upper Limits
tion is a measure of the prospects for FeO, and thus the Some of the calculated oxygen fugacities for parent
∆IW expected for the body taken as a whole, including bodies accreting onto WDs are constrained as either me-
accreted rock and ices. dians with upper bounds, or as upper limits in this work.
We follow Doyle et al. (2019) and assign the mole These cases arise when there is a dearth of oxygen and
fraction of Fe in the metal core (xmetal
Fe ) to 0.85 (Mc- the calculated mole fraction of FeO is small in compari-
Donough 2003). Variations in ∆IW of less than 0.2 dex son to the errors (Doyle et al. 2019). In these cases, the
can occur where xmetal
Fe varies from 0.80 to 1.0, consis- abundance of Fe observed in the WD represents accre-
tent with estimated ranges of bodies in the Solar System tion of a significant amount of Fe metal, and errors in
and constraints on core-like objects polluting WDs (e.g. all of the elemental abundances accumulate in the calcu-
Hollands et al. 2018). lation of the value for xrock
FeO (Equation 4). These errors
Where are the Extrasolar Mercuries? 5

1.0 1.0

WD 1551+175 Ton 345


Doyle et. al (2019)
0.8 0.8
Including negative values
Cumulative frequency

Cumulative frequency
0.6 0.6

ΔIW = -2.098
0.4 0.4

ΔIW = -0.847
0.2 ΔIW = -1.408 0.2

ΔIW = -1.397

0.0 0.0
-1.0 -0.5 0.0 0.5 1.0 -1.0 -0.5 0.0 0.5 1.0
XFeO XFeO

Figure 2. Cumulative frequency distributions for calculated mole fractions of FeO (xrock FeO ) for two polluted white dwarfs. On
the left are results for WD 1551+175. A well defined xrock
FeO is obtained for this case, yielding a well-defined ∆IW value. Including
negative values for xrock
FeO in our Monte Carlo bootstrap method (red) has little effect on the median value for ∆IW obtained
compared to the value where only positive values are accepted in the Monte Carlo draws for xrock FeO (blue). On the right are results
for WD Ton 345. Including negative values for xrockFeO (red) decreases the median by about 0.25 relative to the value calculated
by excluding negative values (blue), resulting in a lowering in the ∆IW value of −1.25 dex compared with the previous method.
These data are sufficient to constrain a median and an upper bound for ∆IW.

8000 In Doyle et al. (2019), we only included Monte Carlo


Lower limit,
G 29-38 median and draws where 0 ≤ xrock rock
FeO . A negative value for xFeO is
GD 40 upper limit ≥ 0
HS 2253+8023 of course non-physical, but arises where errors are large
6000
in comparison to elemental abundances. Negative mole
fractions of FeO imply no FeO in the rock component
number

upper limit and


median ≥ 0
4000
of the parent body because there are more metals than
can be accounted for by oxygen (i.e. Oxs (Equation 4) is
less than Fe). In these instances, oxygen fugacity is or-
2000 ders of magnitude lower than that of the Iron-Wüstite
upper limit ≥ 0
buffer. For the majority of WDs in this study, 11 of
16, random draws infrequently yield negative values for
0 xrock
FeO and eliminating these negative values does not sig-
-3 -2 -1 0 1
xFeO nificantly affect calculated median ∆IW values. For ex-
ample, as shown in Figure 2, the median ∆IW value
Figure 3. Comparison of parent body FeO mole fractions for WD 1551+175 changes by < 0.01 dex if we include
obtained by Monte Carlo error propagation from WDs that
draws where negative values for xrockFeO are obtained, as
yield an upper limit (red), a median with an upper bound
but no lower bound (yellow), and a median with upper and opposed to redrawing those values.
lower bounds (blue) for xrock rock
FeO . We assign upper limits to xFeO
However, for five WDs in this study, lower bounds
and the corresponding ∆IW values where the Monte Carlo on ∆IW cannot be obtained because the lower bounds
bootstrap method yields a proportion of negative values for on xrock
FeO , calculated as the 16.5 percentile in the Monte
xrock
FeO ≥ 50%. Where the Monte Carlo bootstrap method Carlo distribution, are negative, precluding calculation
yields negative xrock
FeO values for between 16.5% and 50% of of a corresponding ∆IW value (Equation 3). For these
the draws, we report a median and an upper bound for ∆IW
we report median values and an upper bound, or, in
derived from the xrock
FeO values.
some cases where the median mole fraction of FeO is
negative, just an upper limit. In the cases where nega-
tive xrock
FeO values are a significant fraction of the Monte
propagate to the calculated relative oxygen fugacities
Carlo draws, we improve upon the method of Doyle et al.
(Equation 3).
(2019) by including xrock
FeO ≤ 0. Excluding large numbers
6 Doyle et al.

Mercury Mercury Earth Vesta


Earth Mars E Chondrite
Mars
C Chondrite
O Chondrite
Extrasolar

E chondrite

SDSS J1043+0855 SDSS J1043+0855


WD 1536+520 WD 1536+520
Solar gas

Solar gas
GD 40 GD 40
SDSS J0738+1835 SDSS J0738+1835
WD 1226+110 WD 1226+110
WD 1145+017 WD 1145+017
WD 1929+012 WD 1929+012
Ton 345 Ton 345
G29-38 G29-38
HS 2253+8023 HS 2253+8023
WD 1350-162 WD 1350-162
WD 1232+563 WD 1232+563
WD 2207+121 WD 2207+121
WD 0446-255 WD 0446-255
G 241-6 G 241-6
WD 1551+175 WD 1551+175

-8 -6 -4 -2 0 -8 -6 -4 -2 0
ΔIW ΔIW

Figure 4. Calculated relative oxygen fugacities for WDs and Solar System bodies in this study. The left panel shows values
obtained using the methods described in Doyle et al. (2019) and the right panel shows values obtained using the methods
outlined in Section 2. Median oxygen fugacities relative to IW are recovered for all but one of the WDs. Solid blue circles are
calculated oxygen fugacities for rocky extrasolar bodies accreting onto WDs in this study where both upper and lower bounds
are obtained. For four of the sources the median value for xrock
FeO is > 0 but the lower bound is not (see Section 2.3). In these
cases medians are shown as open circles. The method used to obtain the values shown on the panel on the right protects against
biasing ∆IW values to higher values and allows us to recover lower medians for SDSS J1043+0855 and Ton 345. Ranges of
∆IW for a gas of solar composition (yellow) and for most rocky bodies in the Solar System (blue) are shown for comparison.
Hypothetical WDs calculated using rocks from Solar System bodies (see Section 2) are also shown for comparison.

of negative random draws from our analysis biases the bound for the derived ∆IW values, updated from Doyle
mole fractions of FeO and therefore the median and/or et al. (2019). The ∆IW values for the body accreted
upper bounds on oxygen fugacities derived from them, by GD 40 is now reported as a median and an upper
to spuriously high values. Including the negative xrock
FeO bound, updated from Doyle et al. (2019), because the
in these cases has a pronounced effect on ∆IW values. frequency of negative draws for xrock
FeO is very close to the
For instance, the calculated median ∆IW for Ton 345 limit we define here (16.5%).
changes by −1.25 dex when negative values are included
(Figure 2). 3. RESULTS
Here we assign upper limits to WDs where ≥ 50%
3.1. Oxygen Fugacity Values
of the the Monte Carlo bootstrap draws (with replace-
ment) yield negative values for xrock
FeO . Where the Monte
Results show that most rocky bodies accreting onto
Carlo bootstrap method yields negative xrock
FeO values for
WDs in this study formed under oxidizing conditions,
between 16.5% and 50% of the draws, we report a me- as did most of the bodies in the Solar System, includ-
dian and an upper bound for ∆IW derived from the ing Earth (Figure 4). Table 1 lists the median values
xrock
FeO values (Figure 3). Using these criteria, we find
for ∆IW and ∼ 1σ errors (from an extension of the in-
one WD for which ∆IW values are constrained only terquartile range to encompass 67% of the distribution)
by an upper limit (G 29-38) and four WDs for which for each WD in this study. Variations in ∆IW of ∼
we report both a median and an upper bound on the 0.2 dex beyond those portrayed in Table 1 can occur
∆IW values (SDSS J1043+0855, WD 1145+017, GD 40, where xmetal
Fe varies from 0.80 to 1.0. The values re-
and Ton 345). Note that new abundance estimates of ported in Table 1 assume xmetal
Fe = 0.85. In Figure 4,
rocky debris accreting onto WD 1145+017 from Fortin- results of WDs using methods from Doyle et al. (2019)
Archambault et al. (2020) yield a median and an upper are compared to results from this study. WDs with ∆IW
values represented by dark blue circles show the results
Where are the Extrasolar Mercuries? 7

Table 1. Oxygen fugacities determined from WD data in this study.

WD Type Temperature (K) ∆IW Oxs /Fe References

SDSS J1043+0855 DAZ 18330 −2.132+1.169 0.337+10.427


−15.209 Melis & Dufour (2017)

WD 1536+520 DBAZ 20800 −1.374+0.284


−0.317 6.280+6.271
−3.705 Farihi et al. (2016)

GD 40 DBZA 15300 −1.262+0.309 2.495+2.629


−2.242 Jura et al. (2012)

SDSS J0738+1835 DBZ 13950 −1.137+0.167


−0.182 10.110+8.577
−5.269 Dufour et al. (2012)

WD 1226+110 DAZ 20900 −0.630+0.307


−0.432 3.360+5.396
−2.210 Gansicke et al. (2012)

WD 1145+017 DBZ 14500 −0.700+0.346 1.215+4.100


−1.650 Fortin-Archambault et al. (2020)

WD 1929+012 DAZ 21200 −0.493+0.255


−0.411 2.023+4.200
−1.486 Gansicke et al. (2012)

Ton 345 DBZA 19780 −2.098+1.544 −0.098+1.428


−1.476 Wilson et al. (2015)

G 29-38 DAZ 11820 < 0.947 −0.183+1.435


−1.413 Xu et al. (2014)

HS 2253+8023 DBAZ 14400 −0.936+0.068


−0.085 2.858+1.151
−1.017 Klein et al. (2011)

WD 1350-162 DZAB 11640 −1.012+0.182


−0.205 4.413+2.533
−1.878 Swan et al. (2019)

WD 1232+563 DZBA 11787 −1.209+0.182


−0.194 15.230+9.915
−6.435 Xu et al. (2019)

WD 2207+121 DBZA 14752 −1.332+0.229


−0.249 7.271+6.574
−4.076 Xu et al. (2019)

WD 0446-255 DZAB 10120 −1.473+0.198


−0.223 4.570+3.719
−2.603 Swan et al. (2019)

G 241-6 DBZ 15300 −1.536+0.266


−0.289 7.368+5.797
−3.791 Jura et al. (2012)

WD 1551+175 DBZA 14756 −1.397+0.178


−0.195 6.377+5.815
−3.705 Xu et al. (2019)

where medians and both upper and lower bounds are ing negative values for xrock
FeO in order to avoid biasing the
obtained. Open circles show instances where medians result. These cumulative frequency distributions pro-
with upper bounds are obtained. In one instance, G 29- vide an immediately discernible indication of instances
38, the median value for xrock
FeO is negative when including where xrock
FeO is evidently low, but where errors are also
all random draws, therefore only an upper limit with no large.
median is reported. We point out that Earth is more reduced than many
Figure 5 shows cumulative distribution functions for bodies in the Solar System, including chondrites, Vesta
all of the WDs in this work. One can see the frequen- and Mars. In order to identify Earth-like extrasolar
cies of negative xrock
FeO values for G 29-38 (DA), SDSS material, we need better ∆IW resolution to distinguish
J1043+0855 (DA), GD 40 (DB), Ton 345 (DB) and WD Earth-like and chondritic values. In implementing this
1145+017 (DB) are greater than for the other WDs. new method and utilizing negative values for xrockFeO , we
These are the five WDs for which lower bounds on the find that the rocky body accreting onto Ton 345 is more
oxygen fugacities are not constrained. In four of these reduced than the previous method would have implied.
five, the frequency of negative draws for xrock
FeO is between SDSS J1043+0855 shows a similar shift to a lower ∆IW
16.5% and 50%, and a positive median is recovered. In value. We reiterate that utilizing negative values for
one case, G 29-38, negative mole fractions are obtained xrock
FeO removes biases that otherwise force relatively low
in more than 50% of the draws, and a positive median ∆IW values towards artificially high values (see Section
xrock
FeO is not recovered, resulting in only an upper limit 2.3), thereby affording more resolution in ∆IW values.
on the value for ∆IW . Inspection of the cumulative fre- The greater resolution enables us to differentiate be-
quency distributions suggests that these five WDs may tween an Earth and a chondritic ∆IW value, for example
be accreting reduced bodies more akin to Mercury. The (Figure 4).
broad distributions underscore the significance of includ-
8 Doyle et al.

1.0 1.0

0.8 0.8
CI chondrite
Ton 345
Cumulative frequency

Cumulative frequency
WD 1145+017
0.6 0.6 GD 40
G241-6
HS 2253+8023
WD 0446-255
0.4 CI chondrite 0.4 WD 1232+563
G29-38 WD 1350-162
SDSS J1043+0855 WD 1536+520
WD 1226+110 WD 1551+175
0.2 WD 1929+012 0.2 WD2207+121
SDSS J0738+1835

0.0 0.0
-1.0 -0.5 0.0 0.5 1.0 -1.0 -0.5 0.0 0.5 1.0
XFeO XFeO

Figure 5. Cumulative frequency distribution functions for the Hydrogen-dominated (left) and Helium-dominated (right) WDs
used in this study. Values for ∆IW are constrained as upper limits where the frequency of negative xrock
FeO is ≥ 50% of total draws
and medians with upper bounds exist for WDs where the frequency of negative draws is such that 16.5% ≤ xrock FeO ≤ 50%. For
these five WDs for which lower bounds are not recoverable, we include negative values in the calculation of the xrock
FeO distribution
in order to avoid biasing the calculated oxygen fugacity. Fictive white dwarfs accreting a debris disk composed of CI chondrite
are shown for comparison in each panel (see text).

Oxs /Fe ratios in Table 1 provide a good indication of the accreting parent body with no corrections for set-
whether a WD will yield a median ∆IW with lower and tling. The veracity of this latter assumption varies from
upper bounds, a median ∆IW with an upper bound but star to star (Koester 2009; Girven et al. 2012).
no lower bound, or just an upper limit on the ∆IW. The The robustness of the derived oxygen fugacities under
value for this ratio is ideally unity for any rock (Equa- these assumptions can be addressed using the effects of
tion 4). Values statistically greater than 1 suggest an- a range of possible scenarios for accretion of material
other source of oxygen, either as the result of differential onto a star together with diffusion of heavy elements
gravitational settling or perhaps accretion of a volatile out of the stellar atmosphere. These two processes can
species (e.g., H2 O). Values less than 1 suggest another interact to affect the observed elemental abundances in
source of iron (metal). Generally speaking, where errors a WD atmosphere. Three different accretion/diffusion
allow Oxs /Fe to be < 1, lower bounds are not recover- phases for planetary debris accreting onto a WD have
able. As values for Oxs /Fe decrease, so does the ability been described previously (Dupuis et al. 1992; Dupuis
to recover this lower bound. Of the 16 WDs reported et al. 1993; Koester 2009) and include an initial build-up
in Table 1, only 2 have negative Oxs /Fe, and the most phase of elements into the atmosphere as the debris ac-
negative value is the same WD that provides only an cretes onto the star, followed by a steady state between
upper limit on ∆IW. accretion onto the WD and settling of elements into the
star, and finally a declining phase as the flux of accre-
3.2. The Effects of Settling on Calculated Oxygen tion decreases and gravitational settling out of the pho-
Fugacities tosphere of the WD dominates. A combination of mass
and ionization state generally leads to heavier elements
We have thus far invoked a correction for differential sinking faster than lighter elements (e.g. iron will sink
settling of the elements assuming a steady-state between faster than oxygen) with some exceptions. Over time,
accretion and diffusive settling for the 4 Hydrogen- these differences in diffusive velocities lead to changes
dominated WDs (DAs) as in Doyle et al. (2019). The in abundance ratios. The effects of settling on observed
assumption of a steady state is generally reasonable in element ratios can be accounted for using models of var-
these cases because of the relatively short characteris- ious levels of complexity.
tic timescales for settling in these WDs. Because of the Extracting element ratios for the accreting parent
comparatively long characteristic timescales for diffusive body from the ratios measured in the WD is most eas-
settling of the heavy elements in helium-dominated WDs ily done when accretion is clearly either in the build-up
(DBs), they are assumed to be in a build-up phase dur- phase or the steady-state phase (Dupuis et al. 1992).
ing which the observed element ratios are attributed to
Where are the Extrasolar Mercuries? 9

In these cases a simple mass balance expression for the parent body, and that the rates of accretion are con-
mass of element z in the convective layer (CV) in terms stant. Equation 10 is the basis for the often applied
of a constant rate of accretion, ṀCV,z , and the rate of steady-state correction to WD element ratios used to
gravitational settling based on a constant diffusive ve- obtain parent body ratios:
locity can be used (Dupuis et al. 1992):
MPB,z1 MCV,z1 τz2
dMCV,z MCV,z = . (11)
= ṀCV,z − (5) MPB,z2 MCV,z2 τz1
dt τz
where τz is the constant e-folding time for diffusive set- In the case of DA white dwarfs, diffusion timescales (τz )
tling through the base of the convective layer. The gen- are on the order of days to thousands of years and are
eral solution is much shorter than typical estimates for the duration of
an episode of accretion represented by a polluted WD
o
MCV,z (t) = MCV,z e−t/τz (τdisk = 104 − 106 yrs) (Girven et al. 2012). In these
Z
(6) cases, a steady-state phase should be reached rapidly,
+ e−t/τz et/τz ṀCV,z (t) dt soon after accretion begins, and we are almost certainly
observing the star post build-up phase. Therefore, for
o
where MCV is the initial mass of z in the convective layer DA white dwarfs, we apply the correction from Equa-
that we generally assume is zero. When settling times tion 11 using calculated values for diffusion timescales.
are long compared with the duration of accretion, t, and Whether the correction should be applied to DB WDs
the rate of accretion is essentially constant, Equation 6 depends on the surface temperature (i.e., τz ) and es-
reduces to timates for accretion rate and accretion duration, the
latter being all but unknown (see Section 4.3).
MCV,z = ṀCV,z t. (7) The inevitable waning of accretion in the declining
One expects this case to be most suitable for a DB WD phase implies that mass accretion rates are not constant
at moderate Teff surrounded by a debris disk since the
settling e-folding times in this case can be of order 105 WDs with debris disks
WDs without known debris disks
to 106 years. The ratio of heavy elements observed in 20 WDs with xFeO within error of 0

the WD is in this case a direct measure of the elemental


ratio in the parent body (PB) assuming that the ratio
Oxs/Fe

of accretion rates accurately reflects the elemental ratio 10


in the parent body. From Equation 7 the relationship
between the ratios of two heavy elements z1 and z2 in
silicate
the WD and in the accreting parent body is therefore 0

MCV,z1 ṀCV,z1 MPB,z1


= = . (8)
MCV,z2 ṀCV,z2 MPB,z2 -10
-6 -4 -2 0 2
Conversely, when the duration of accretion is relatively ΔIW
long compared with the e-fold time for settling in the
Figure 6. ∆IW vs oxygen in excess of that needed to oxidize
WD, t >> τz and e−t/τz << 1, and the solution to
Mg, Si, Ca, and Al (Oxs ), relative to Fe. DA and DB white
Equation 6 becomes dwarfs with confirmed debris disks are shown as dark blue
circles. Light blue squares represent WDs without confirmed
MCV,z (t) =ṀCV,z τz (1 − e−t/τz ) debris disks. WDs yielding medians with upper bounds, but
(9)
=ṀCV,z τz . no lower bounds for ∆IW are shown with open gray sym-
bols. These limits come about when, within errors, there is
Under these circumstances, the ratio of the masses of a dearth of oxygen and FeO is not present in a large fraction
two elements in the WD convective layer is related to of the random draws (Doyle et al. 2019). The horizontal dot-
that in the accreting parent body by ted line represents silicate and oxide rock that, in principle,
would always lead to Oxs /Fe = 1, assuming the amount of
MCV,z1 ṀCV,z1 τz1 MPB,z1 τz1 Fe3+ is small compared with Fe2+ . WD 1232+563 does not
= = (10) have a confirmed debris disk and has the highest value for
MCV,z2 ṀCV,z2 τz2 MPB,z2 τz2 Oxs /Fe in this study, meaning a declining phase of accretion
where again it is assumed that the accretion rate ra- may be a reasonable inference for this WD.
tio faithfully reflects the elemental ratio of the accreting
10 Doyle et al.

WD 1226+110-like O, taudisk = 1E5


Si
GD 40-like
1018
Fe
Mg
1022
Al
Ca
1016

Moles in mixed layer


Moles in mixed layer

1020

1014

0.0 0.0
1018

-1.0
1012
ΔIW

ΔIW
-1.0
-2.0
1016
-2.0 -3.0
1010 100 102 104 106 100 102 104 106
time (yrs) time (yrs)

1014
100 101 102 103 104 105 106 100 101 102 103 104 105 106
time (yrs) time (yrs)

Figure 7. Model for the time-dependent element accretion history onto a DA similar to WD 1226+110, left, and a DB similar
to GD 40, right, using the model for an exponential decay of a debris disk from Equation 13, assuming a debris disk e-folding
lifetime of 1 ×105 years. Diffusion timescales are from Koester (2013). The insets show variations in ∆IW over this timescale.
The parent body ∆IW values are indicated as horizontal dashed lines in the insets. The fictive WD 1226+110 initiates with
molar compositions corrected using a steady-state approximation (Equation 11) representative of the parent body. For the DA
WD, ∆IW is robust throughout all accretion/diffusion phases, because diffusion timescales are short and thus relative diffusion
timescales are small. For the DB WD, if the accretion event is longer than ∼ 2 ×106 years, one could see a spurious ∆IW value
indicative of reduced material (∆IW < −3) compared with the parent body value of ∼ −1.

(as assumed in Equation 5). As the diffusion of elements Also shown in Figure 6 is the effect of a paucity of O in
out of a WD atmosphere commences and the build-up a WD in causing an apparent abundance of iron metal,
phase ends, a relative increase in the abundance of oxy- which in turn results in an indeterminate lower bound
gen, O, and/or Mg will occur as differences in diffusion for ∆IW rather than a unique determination. The up-
velocities cause Si, Al, Ca, and Fe to sink faster. This per limit arises in the case of apparent metal because
causes the observed excess oxygen to rise (Equation 4). in our calculation, xrock
FeO is constrained by the smaller
Most WDs in this study have oxygen excesses beyond of the two abundances, Oxs or Fe. Thus, if a parent
that needed to form FeO. It is important to note that, body has a large bulk Fe content, or a relatively small
in these cases, xrock
FeO is limited by the abundance of Fe, bulk O content, xrock
FeO will be equivalent to Oxs . From
not oxygen. This is shown in Figure 6 where the or- Equation 4 it is clear that the errors associated with
dinate, Oxs /Fe, is > 1. In most cases, element ratios excess oxygen are accumulated from those of all of the
suggest that these excesses are indicative of the accre- major rock-forming elements, and are thus considerably
tion of rocky material, including crust (basaltic) and/or greater than those for Fe alone, as described in detail
mantle without a metallic core. In some cases the oxy- by Doyle et al. (2019). These large errors result in the
gen excesses are much larger than can be accounted for indeterminate lower bounds on ∆IW values described
by accretion of ices (e.g. H2 O, CO) with rock. In these above.
extreme cases, a large value for Oxs may be the result The potential importance of settling emphasizes the
of accretion in a post build-up or declining phase. The need to consider the effects of variable accretion rates
existence of debris disks around most of the WDs in on element ratios in the mixing layer of the polluted
this study decreases the likelihood that we are observ- WDs. Jura et al. (2009) formulated a model for the
ing stars in the declining phase, in which the material time-dependent mass of an element z in the convective
in the WD atmosphere is residual, because there is still layer, MCV,z (t), in terms of an evolving mass accretion
material in a disk that has yet to be accreted. However, rate. For simplicity, they posited that the accretion rate
for WDs such as WD 1232+563, which lacks a debris onto the WD is a function of the mass in a circumstel-
disk and has the highest Oxs /Fe ratio in Figure 6, a lar accretion disk, and that the e-folding time for the
declining phase of accretion is a reasonable inference. declining mass of the disk attributable to viscous dissi-
Where are the Extrasolar Mercuries? 11

pation, τdisk , is a constant. This results in an accretion rather than oxidized material (∆IW < −3). On the
rate that decreases exponentially with time. The equa- other hand, the model for the time evolution of ∆IW in
tion that replaces Equation 5 under these circumstances the model resembling WD 1226+110 reveals that ∆IW
is is an extremely robust parameter. This is because diffu-
sion timescales in DA WDs are short, limiting the rela-
dMCV,z o
MPB,z e−t/τdisk MCV,z tive diffusive separation of the elements. In what follows
= − , (12)
dt τdisk τz we concentrate on DB white dwarfs because the oxygen
o
where MPB,z is the initial mass of z in the parent body fugacities calculated from these WDs are evidently sen-
that forms the circumstellar disk (and the initial mass sitive to the duration of accretion and diffusion at the
of z in the disk). The solution (see Appendix) is time of the observations.

4. DISCUSSION
o
MPB,z τz h −t/τdisk i
MCV,z (t) = e − e−t/τz . (13) 4.1. Accretion of an Oxidized Parent Body
τdisk − τz

Koester (2009) also examined the effects of Equation 13 PB value


for element ratios in a polluted WD. Here we use this
0
model to explore the consequences of elemental settling -8 -6
ΔIW
-4 -2

through the WD envelope on calculated oxygen fugaci-


ties as a function of time. For the purposes of this work, 20,000

in the first instance we assume that τdisk = 105 yrs. The


reasonableness of this assumption is discussed in Section 18,000

4.3.
Temperature (K)

Figure 7 shows the calculated evolution of moles of 16,000

elements in the mixing layer of a WD with time, based


14,000
on Equation 13, using the examples of a DA similar
to WD 1226+110 (Teff ∼ 21000K), and a DB similar
12,000
to GD 40 (Teff ∼ 15000K) (Gansicke et al. 2012; Jura
et al. 2012). In both cases, the parent body oxygen fu-
10,000
gacities, expressed as ∆IW values, are taken to be near
−1, representing relatively oxidized materials in the con- 0 5.0x105 1.0x106 1.5x106 2.0x106
time (yrs)
text of solar system values (Doyle et al. 2019). For this
calculation, we used diffusion timescales from Koester Figure 8. Evolution of calculated ∆IW values for a model
(2013) based on log g = 8.0 cm s−2 . For the elements WD accreting a parent body (PB) composed of CI chondrite
accreting onto the model similar to GD 40, one can see material (∆IW = −1) as a function of both time and effec-
the build up of heavy elements in the atmosphere, fol- tive temperature for the WD. The accretion/diffusion model
lowed by an interval of steady state (at the maximum is that for an exponentially decaying debris disk as in Equa-
tion 13, assuming a disk lifetime of 1 × 105 years. Hotter
of the curves) and finally a period in which the heavy
colors are more oxidized ∆IW values and cooler colors are
elements settle into the star as accretion from the disk less oxidized values. Diffusion timescales are interpolated as
wanes. In the model DA white dwarf (left), short dif- a function of Teff from Koester (2013) based on log g = 8.0
fusion timescales result in a rapid approach to steady cm s−2 . As done with helium-dominated WDs in this study,
state, as expected, effectively precluding an observation elemental abundance ratios with no steady-state corrections
of the build-up phase. In both cases, once accretion (i.e., Equation 8) were used to calculate ∆IW values.
from the circumstellar disk begins to decrease signifi-
cantly, element ratios change markedly, beyond where
a simple steady-state correction can compensate. The Figure 7 shows the evolution of material accreting
∆IW values for the model GD 40 decrease rapidly once onto two particular WDs. In Figure 8 we generalize
accretion has decreased enough that settling dominates, the time-dependent model to consider an exponentially
causing Fe to decrease in abundance relative to other el- decreasing disk of relatively oxidized CI chondrite mate-
ements, thereby decreasing xrock
FeO and ∆IW. For this star, rial accreting onto helium-dominated white dwarfs vary-
under these assumed conditions, if the accretion event is ing in temperature from 9000 to 21000K. This range in
longer than ∼ 2 ×106 years, the calculated ∆IW value Teff encompasses all of the He-rich white dwarfs in this
would be spuriously low, suggestive of reduced material study. We interpolate diffusion timescales as a function
12 Doyle et al.

100% 2 Al
Al Ca
Si
90%
Ca Mg
Fe
O
80% Si 0

steady state

declining
build-up
70% Mg

60%
Fe -2
O
Mass

ΔIW
50%

40% -4

30%

20% -6

10%

0% -8
ar th y te B 0 5x105 1x106 1.5x106 2x106 2.5x106 3x106
l ar ur dri P
So lk E erc on arth
Lh MO y
zo B
e
E
M th
Ve n
a
er rs
st ry
ite
L
W 04 Ea I
15 0 h

W 3 D 0

+ 0
7
+ 5

W 12 +1 0
11 + 5

time (yrs)
C

M Ch E
lit

st
lk BS

oo

D 3+ rt

52

01
36 85

45 11
D 6 3
er R

Bu
M a
En cu
r

at

I d
Ea

C ce
G

u
ed
J1 ced

D 8
2

R
Bu

7
SS du

J0
SD Re

SS

Figure 9. Reduced Earth model depicting variations in abundance ratios with time in the mixing layer of a polluted WD due to
SD

differences in diffusion fluxes out of the bottom of the layer at Teff ∼ 15000K. The time-variable accretion and diffusion model is
that defined in Equation 13, and includes a debris disk that decays exponentially with an e-folding time of 1 × 105 yrs. Diffusion
timescales are from Koester (2013) and are based on log g = 8.0 cm s−2 . The colors represent the normalized abundances of
the six major rock-forming elements at each time step in the model. The black line represents the calculated value for ∆IW at
each time step. The apex of the ∆IW curve represents the time at which xrock FeO changes from being constrained by Oxs to being
constrained by the abundance of Fe. The exact locations of the vertical dotted lines representing shifts in accretion phases are
discussed in the text. The reduced Earth model yields ∆IW −5.9, using our standard method in Section 2, and is indicated
by the horizontal dashed line. The relative elemental abundances for Solar System bodies and the reduced Earth parent body
(PB) are shown for comparison at left.

of Teff again from values calculated by Koester (2013) folding time for the debris disk of 105 yrs, the result
and based on log g = 8.0 cm s−2 . For the purposes of this would be an incorrect conclusion that the parent body
o
calculation, we assume the total initial mass, MPB , is rock was relatively reduced rather than oxidized. This
equivalent to the mass of asteroid 1 Ceres. For any given is the same phenomenon as seen in the fictive GD 40 in
temperature, one sees variations in ∆IW throughout the Figure 7. In this study, there are no WDs that exhibit
various accretion/diffusion phases. For the most part, such low ∆IW values (< −3), suggesting that the DBs
the model suggests that when oxidized CI chondrites with Teff in this range are not accreting material in a
actively accrete onto a DB white dwarf, ∆IW should declining phase millions of years after the start of the
vary with time by less than 0.5 dex from the parent body accretion event.
∆IW value of ∼ −1. However, for DB white dwarfs with
effective temperatures of approximately 16000−19000K, 4.2. Accretion of a Reduced Parent Body
calculated oxygen fugacities for the parent body could
Of the 16 WDs assessed thus far, 11 are consistent
decrease to spurious values with ∆IW < −3 during the
with the accretion of oxidized (∆IW > −3) rocky bod-
declining phase as a result of gravitational settling of
ies, while 5 may be accreting more reduced material. In
elements out of the white dwarf atmospheres. For the
the Solar System, one terrestrial planet, Mercury, is re-
assumptions used in these calculations, this phase would
duced, as are a small percentage of meteorites (Figure
begin after about 5×105 yrs, after 5 disk e-folding times.
4). If the Solar System were a guide, one might expect to
At these Teff , if a WD polluted by CI-like material were
see a Mercury-like oxygen fugacity in ∼ 3% of polluted
observed after a million years of accretion with an e-
WDs. Here we evaluate the conditions under which the
Where are the Extrasolar Mercuries? 13

very low fO2 values like those exhibited by Mercury, en- In the case of the reduced Earth model, initial xrockFeO is
statite chondrites, and aubrites (the igneous equivalent < 0.001.
of enstatite chondrites) could be retrieved from a pol- The intervals for the different accretion/diffusion
luted WD using the current method. phases in Figure 9 are indicated by the dashed, vertical
Consider the case in which a white dwarf is accreting lines. The build-up phase lasts < 3 × 105 years, during
debris from a disk composed of Mercury, with a large which time abundances observed in the WD atmosphere
iron core and low FeO in the rock. If we were to assume are similar to parent body abundances. The apparent
that the material is accreting onto the star in a build-up oxygen fugacity increases dramatically even during this
phase by equating parent body abundance ratios with stage. We regard the end of the build-up phase as the
observed abundance ratios of the heavy elements, but time at which dXO /dt = 0 (Figure 10). We consider the
the material is actually accreting in a steady-state phase, onset of the declining phase to be 5 disk e-folding times,
the high relative abundance of O resulting from the un- or 5 × 105 years into the model, which corresponds to
accounted for settling would cause us to overestimate the time at which the debris disk is less than 1% of its
xrock
FeO and the intrinsic oxygen fugacity for the parent initial mass. The influence of the decrease in mass of the
body. This is because there is plenty of iron that is avail- debris disk, and our definition of the declining phase, is
able to pair with the oxygen that has been enhanced by clear when the same model as that shown in Figure 9 is
differential gravitational settling, causing a seemingly run with no gravitational settling (Figure 10).
high FeO concentration by our calculation whereas in The discontinuity at the apex of the ∆IW curve oc-
reality the iron existed as metal. curs at about the initiation of the declining phase at
In order to investigate further the evolution of ∆IW which xrock
FeO is no longer constrained by Oxs (i.e., where
for a reduced body, like Mercury, accreting onto a DB moles Fe ≥ moles O) but rather by the abundance of
white dwarf, we use a model for a reduced Earth anal- Fe in the WD (i.e. moles Fe ≤ O; Figure 9). There-
ogous to an enstatite chondrite model for Earth; we do fore, if a reduced bulk Earth were to accrete onto this
not use Mercury owing to the likely anomalous size of its hypothetical DB, its ∆IW value would be constrained
core (Chau et al. 2018). Our reduced Earth model has as an upper limit in our calculation prior to the de-
the bulk composition and mass of Earth, but the 8 wt. clining phase, suggestive of the reduced nature of the
% FeO in the rock is decreased to <1 wt. % by removing parent body. However, it would no longer be an upper
the necessary amount of oxygen. The Fe liberated from limit once the declining phase progressed for 1.5 × 106
the oxide is assigned to the metal core, increasing the years. Rather, a moderately oxidized value for ∆IW
core mass from 32 wt. % to 37 wt. %. would be obtained until about 3 × 106 yrs of settling,
Figure 9 shows the time-dependent model for an ex- and there would be no signature of the low oxidation
ponentially decaying disk composed of debris from the state of the parent body during this interval. The like-
reduced Earth accreting onto a DB white dwarf where lihood for completely disguising a reduced Earth is high
Teff = 15000K. The colors represent the normalized con- once the pollution episode extends to approximately 5
centrations of the rock-forming elements and the black disk e-folding times or greater in this model, after which
curve shows the calculated ∆IW values where ∆IW point there is no longer evidence that metal existed in
= −5.9 at t = 0. One can see that in this model the parent body. Similarly, if individual accretion events
the reduced body with an abundance of iron metal ap- last for < 5 × 105 years in the context of this model (i.e.,
pears oxidized late in the accretion event. The observed < 5 e-folding times for the disk), our calculation method
composition of the body with no accounting for settling for ∆IW should be capable of distinguishing potentially
seemingly evolves from a mixture of metallic core and reduced parent bodies from oxidized bodies, even if only
rock to a rock-like material anomalously rich in Mg af- by the mere existence of an upper limit constraint and a
ter 106 yrs. We note that this composition influenced by high propensity for negative mole fractions of FeO. The
settling does not resemble the compositions in Figure 1. uncertainty of disk lifetimes is discussed in Section 4.3.
After 100,000 years into the model, the relative mass of Generalizing this result, when the disk lifetime is com-
O has increased enough to alter the calculated oxygen parable to typical settling times through the mixed layer
fugacity by approximately +2 dex, but the oxygen fu- of the polluted WD, and observations are taken after ap-
gacity would still appear to be relatively low at ∼ −4. proximately 5 disk e-folding times or more have passed,
The rapid rise in calculated ∆IW in the beginning of the calculated oxygen fugacities will be biased against find-
accretion episode is because small differences in xrock
FeO are ing reduced parent bodies. In this study, the existence
more important when the parent body is composed of of observable debris disks around most stars decreases
reduced material where initial values for xrock
FeO are small. the likelihood that we are observing the WDs after sev-
14 Doyle et al.

1028 1028

O, taudisk = 1E5
Si
1026 Fe 1026
Mg
Al
Moles in mixed layer

Moles in mixed layer


Ca
1024 1024

1022 1022

1020 1020

1018 1018
100 101 102 103 104 105 106 100 101 102 103 104 105 106
time (yrs) time (yrs)

Figure 10. Molar abundances of rock-forming elements in the atmosphere of the same hypothetical WD as that shown in
Figure 9. Parent body ratios are the values at time zero. In order to illustrate the build-up phase, we are using log time
(unlike Figure 9). The panel on the left represents the time-dependent accretion history of the hypothetical WD as presented in
Equation 13. The panel on the right represents the model, but with no gravitational settling through the WD envelope. Note
that observable abundances scale with the y-axis. Evaluation of the derivative dXO /dt in the left panel gives the calculated
time at which the build-up phase concludes, ∼ 3 ×105 years. Using the figure on the right, the heavy element concentrations
reach their final values after 3 to 5 × 105 yrs, which represents the onset of the declining phase during which the debris disk is
depleted.

eral disk e-fold times unless the disks were considerably rect for settling combined with waning accretion. This
more massive than is customarily inferred from masses occurs at times greater than about 5 × 105 yrs in Figure
in the mixing layer and estimates of accretion rates (see 11. Below ∼ 10000K, diffusion timescales are longer,
Section 4.3). Thus, the probability of masking a reduced closer to the assumed disk lifetime. Therefore, the cal-
body is commensurately low. culated value for ∆IW remains low for the duration of
In Figure 11 we show ∆IW values obtained by ex- the accretion event well beyond the practical lifespan of
panding the model for accretion of a reduced Earth the disk (Figure 11). Slow diffusion of elements through
shown in Figure 9 to include stars varying in Teff from the white dwarf atmosphere means the relative flotation
9000 − 21000K. As in Figure 8, we interpolate derived of oxygen will never dominate over accretion in these
diffusion timescales for each Teff in DBs from Koester cases. In this case, elements do not begin to settle out
(2013) and invoke a characteristic disk lifetime of 1×105 of sight until the majority of the parent body has been
yrs. The variations in ∆IW values at T∼15000K are accreted, and the apparent oxygen fugacity decreases
equivalent to the black curve in Figure 9. because iron is the monitor of ∆IW.
Figure 11 shows that low ∆IW values are for the most Based on our models, the two DBs with Teff > 19000K
part recoverable for DB WDs with Teff < 19000K less in this study (Ton 345 and WD 1536+520) would always
than 1 × 106 years into the accretion event (i.e., for appear oxidized if they were accreting in a declining disk
ten disk e-folding times). In DBs hotter than approxi- phase in which a reduced Earth-like body accretes with
mately 19000K, diffusion timescales become so fast that τdisk similar to settling times. However, the data indi-
DB white dwarfs begin to resemble DA white dwarfs in cate instead that these parent bodies may be reduced
that after an initial episode of oxygen enrichment they rather than oxidized. Ton 345 yields a median parent
rapidly approach a steady state. This is the worst case body ∆IW of −2.1. This value is lower than could be
scenario for detecting low oxygen fugacity parent bod- preserved due to settling unless we were observing the
ies; for a DB WD with Teff > 19000K, preserving low accretion in the first 103 years of the event. Ton 345 does
values for ∆IW is not possible beyond the steady-state possess a gas emission disk, which may suggest an early
phase. This is true even when using Equation 11, as is stage of accretion, as these disks may come from col-
routinely done for DAs, because once the system leaves liding fragments of planetesimals (Manser et al. 2020).
the steady-state phase, there is no a priori way to cor- For WD 1536+520, if one were to calculate ∆IW using
Where are the Extrasolar Mercuries? 15

PB value PB value

-8 -6 -4 -2 0 -8 -6 -4 -2 0
ΔIW ΔIW

20,000 20,000

18,000 18,000

Temperature (K)
Temperature (K)

16,000 16,000

14,000 14,000

12,000 12,000

10,000 10,000

0 5.0x105 1.0x106 1.5x106 2.0x106 0 5.0x105 1.0x106 1.5x106 2.0x106


time (yrs) time (yrs)

Figure 11. Evolution of calculated ∆IW values for model Figure 12. Evolution of calculated ∆IW values as a func-
WDs accreting reduced Earth-like material (see text) as a tion of time and Teff for model WDs accreting reduced Earth-
function of time and Teff . The accretion/diffusion model like material like that shown in Figure 11, but with a smaller
is that for an exponentially decaying debris disk expressed core mass of 3.7 wt. %. Diffusion timescales are interpolated
in Equation 13, assuming a disk e-folding lifetime of 1 × as a function of Teff from Koester (2013) and are based on
105 years. It is the same model as shown in Figure 9, but log g = 8.0 cm s−2 . The ∆IW value for the model parent
expanded to include a range of WD temperatures onto which body (PB) is −5.9 and the most oxidized value calculated for
the parent body accretes. The ∆IW value for the model these fictive WDs is ∆IW −2.86. With such a small amount
parent body (PB) is −5.9, as indicated in the legend. Hotter of iron metal, there is not enough to pair with floating oxy-
colors are more oxidized calculated ∆IW values and cooler gen to significantly alter xrock
FeO and mask the reduced nature
colors are less oxidized. Diffusion timescales are interpolated of the parent body. As done with helium-dominated WDs in
as a function of Teff from Koester (2013) and are based on this study, elemental abundance ratios with no steady-state
log g = 8.0 cm s−2 . As done with helium-dominated WDs in corrections (i.e., Equation 8) were used to calculate ∆IW
this study, elemental abundance ratios with no steady-state values.
corrections (i.e., Equation 8) were used to calculate ∆IW
values.
ward, identification of reduced bodies requires a more
nuanced approach. For Teff < 18000K we should expect
a steady-state assumption, as considered likely by Farihi to see evidence of low oxidation states in the data for
et al. (2016), one gets a median with an upper bound even long accretion times relative to the lifespan of the
for ∆IW, contrary to the unique determination reported debris disk. Hints that the WD is in the waning stages
in Table 1 (see discussion surrounding Figure 13). Here of accretion may show up as odd elemental abundances
again, evidence for low xrock
FeO values at high Teff suggests compared with expectations for most rocks, including
a short duration of accretion. those evidenced in the polluted WDs (e.g., Mg-rich com-
Therefore, when iron metal is present and available to positions). Nonetheless, large amounts of metal will al-
“pair” with floating Oxs using our calculation method ways result in large errors for calculated ∆IW values.
in a scenario like that modeled here, it is possible for a Fortunately, many polluted WDs show evidence of ac-
reduced parent body to appear as oxidized in the WD cretion of mainly rock, perhaps because rock and metal
solely due to differences in diffusion timescales. For our are separated during the accretion process (e.g. Manser
model reduced Earth with a ∆IW value of −5.9, we find et al. 2019). The WDs yielding upper limits in ∆IW as
that in order for the calculated ∆IW value to signify a the result of an apparent abundance of iron metal are
reduced parent body with ∆IW ∼ −3 or less, irrespec- at present the most likely candidates for reduced mate-
tive of time of the observation and at any temperature, rial (low ∆IW values) and should be the focus of future
the object would need to be composed of < 4 wt. % iron investigation.
metal (Figure 12).
Using these models as guides, one concludes that al-
though identification of oxidized bodies is straightfor- 4.3. Uncertainties in the Calculations
16 Doyle et al.

The model shown in Figure 11 makes clear that crucial Large ranges in accretion rates may imply that accre-
parameters necessary to evaluate the veracity of calcu- tion is indeed episodic, as has been suggested by others
lated ∆IW values for DB WDs are the duration of ac- (e.g. Farihi et al. 2012; Raddi et al. 2015; Jura & Xu
cretion prior to the observations and the characteristic 2012). Additionally, the assumption that the accreting
timescale for depletion of the debris disk. These can material is well-mixed, and therefore representative of
not be estimated independently at present and are con- the bulk composition of the parent body may be too
strained only with certain assumptions. One can calcu- simplistic (e.g. Xu et al. 2019). In the future, accretion
late minimum disk lifetimes using the measured masses rates and disk lifetimes will need to be better determined
of all of the rock-forming elements (Z, total) in the for individual WDs in order to better quantify the relia-
helium-dominated WD convective layer (MCV,Z ,total ) bility of geochemical parameters deduced from polluted
with calculated accretion rates (ṀCV,Z ,total ). Pairing WDs.
typical steady-state accretion rates with typical lower
limits for parent body masses as evidenced by the masses
of heavy elements in the convective layers allows for an Mercury Earth Vesta
Mars C Chondrite
estimate of minimum disk lifetimes based on O Chondrite
Extrasolar (build up)
Extrasolar (steady state)

MCV,Z ,total E chondrite


τdisk = . (14)
ṀCV,Z ,total SDSS J1043+0855 (DA)
WD 1536+520

Solar gas
GD 40
SDSS J0738+1835
Using a range for ṀCV,Z ,total of 108 to 1011 g s−1 based WD 1226+110 (DA)
WD 1145+017
WD 1929+012 (DA)
on theory and observations (e.g. Rafikov 2011a,b; Far- Ton 345
G29-38 (DA)
ihi et al. 2012; Wyatt et al. 2014; Xu et al. 2019) and HS 2253+8023
WD 1350-162
the average masses of rock-forming elements observed WD 1232+563
WD 2207+121
WD 0446-255
in the convective zones of the DBs in this study one ob- G 241-6
WD 1551+175
tains minimum disk lifetimes of 4.5 × 104 to 4.5 × 107
years. Using estimates for the mass of rock-forming el- -8 -6 -4 -2 0
ΔIW
ements in polluted DBs from the literature, the same
calculation yields minimum disk lifetimes ranging from Figure 13. Comparison of calculated ∆IW values assuming
3 × 104 to 5 × 106 years (Girven et al. 2012). These that the DBs in this study are accreting in the build-up phase
ranges include the nominal value of 1 × 105 years that vs. values obtained by assuming they are in steady state be-
we used in our modeling. We emphasize, however, that tween accretion and gravitational settling. Values assuming
these estimates rely on the notion that accretion rates a build-up phase are represented by unfilled or filled navy
are constant. Episodic accretion is not captured by this circles, depending on whether ∆IW is a median with an up-
per bound (unfilled) or a median with both lower and upper
calculation.
bounds (filled). Values for ∆IW for WDs assumed to be
The contours in Figures 11 will vary as a function of in a steady-state phase of accretion are represented in the
the values for τdisk . The diffusion times, τz , for the sub- same way, but with red circles. The ability to calculate the
set of DBs in this work range from 103 to 106 years. lower bound for WD 1536+520 is lost when steady-state is
Longer disk lifetimes where τdisk >> τz lead to more re- assumed. DA white dwarfs are indicated by their labels and
gions in Teff vs. time space where the oxidation states of are generally considered to be in a steady-state phase of ac-
highly reduced bodies would be overestimated. Shorter cretion. Here we show G 29-38 assuming both a steady-state
and a build-up phase of accretion, in consideration of the
disk lifetimes where τdisk << τz lead to greater fidelity
possibility of thermohaline instabilities in this DA WD (see
in recording the oxidation states of reduced bodies. Our text). G 29-38 provides an upper limit for ∆IW under a
assumed disk characteristic lifetime of 1 × 105 years is steady-state assumption, but we can define the median un-
a plausible median value, but it is unlikely to be uni- der the assumption of a build-up phase. The differences in
versally applicable. Higher accretion rates, by several ∆IW between steady-state and build-up assumptions for the
orders of magnitude, are obtained when considering the other DAs are given in the text. Generally, differences in
effects of thermohaline mixing and convective overshoot ∆IW when WDs are considered to be in steady state vs. a
(e.g. Deal et al. 2013; Bauer & Bildsten 2018, 2019; build-up phase are less than 0.5 dex.
Tremblay et al. 2017; Kupka et al. 2018; Cunningham
et al. 2019). Higher accretion rates lead to shorter mini-
mum disk lifetimes for the same minimum masses (Equa- Our ability to detect low ∆IW values depends on our
tion 14). ability to assess whether a WD is in a build-up, steady-
Where are the Extrasolar Mercuries? 17

state, or declining phase of accretion. These assessments lead to increases in the initial intrinsic oxidation state of
can be highly uncertain. Thus, Figure 13 shows the cal- the rock. These multiple paths to oxidation have been
culated values for ∆IW considering both scenarios for well-studied through modeling efforts, though it is not
DBs, a build-up and a steady-state phase. Generally known which process is dominantly responsible for ox-
speaking, for the DB WDs, the median values for ∆IW idizing bodies in the Solar System (Krot et al. 2000;
change by <0.5 dex when DBs are assumed to be accret- Grossman et al. 2012; Monteux et al. 2018; Rubie et al.
ing rocky material in a steady state, rather than in the 2015).
build-up phase that we have assumed previously (Figure Most of the rocky bodies accreting onto the WDs in
13). However, in one case (WD 1536+520) assuming a this study are consistent with oxidizing conditions at the
steady-state phase results in the loss of the lower bound time and place of parent body formation. This could
on the median ∆IW value of −0.113 compared to the imply that there is something about rock formation in
value in Table 1 of −1.374, this is the largest shift in general that may favor oxidation relative to a solar gas
median ∆IW values calculated. and that the formation of planets with Earth-like oxygen
For the DA white dwarfs for which element ratios are fugacities may be more common than forming planets
usually corrected assuming a steady-state phase of ac- with Mercury-like oxygen fugacities.
cretion, modeling suggests that some of them are can- However, the difficulty in detecting and accurately
didates for thermohaline instabilities (Deal et al. 2013; measuring the oxidation states of reduced bodies could
Bauer & Bildsten 2018, 2019). If thermohaline instabil- mean that more reduced bodies exist and we have not
ities are present, measured element ratios should not be yet detected them. A relationship between oxidation
corrected with the short settling times typically used. As state of rock and distance from the Sun has been sug-
an illustration, in Figure 13 we show results for DA G 29- gested for the Solar System (Wade & Wood 2005; Wood
38 with and without settling time corrections. Element et al. 2006; Rubie et al. 2011, 2015; Badro et al. 2015).
ratios in G 29-38 do not define a median ∆IW value If this relationship where reduced bodies form prefer-
when considered to be in steady state, but do define a entially closer to the Sun is typical, then it could be
median when a build-up phase is assumed (i.e., no cor- that reduced bodies are destroyed by their host star
rections are applied). Under the assumption of a build- when the star is on the red giant branch (e.g. Schrder &
up phase due to thermohaline instabilities, two of the Connon Smith 2008). This would further decrease the
DAs in this study (WD 1226+110 and WD 1929+012) likelihood of discovering a reduced body by the method
produce ∆IW medians of −1.616 and −1.948, respec- detailed in this paper.
tively, and upper bounds of −0.696 and −0.673, respec-
tively. The median ∆IW for SDSS J1043+0855 is lost 5. CONCLUSIONS
when a build-up phase is assumed, but the upper limit
The utilization of polluted WD stars to obtain detailed
decreases to ∆IW = −2.089. Given the plenitude of
geochemical characterizations of extrasolar rocky bodies
WDs in this study with debris disks, accretion of mate-
is a powerful tool for elucidating the internal composi-
rial in a declining phase is less likely, with the possible
tions and structures of terrestrial exoplanets. This study
exception of WD 1232+563, as discussed in Section 3.2.
analyzes observations of 16 white dwarf stars to under-
stand the oxidation states of the planetary systems at
4.4. Implications for Oxidation States in Planetary the time and place that the parent bodies of the ac-
Science creted material formed. We find that most extrasolar
The process that gave rise to the oxidation of most rocky bodies formed under oxidizing conditions, com-
rocky bodies in the Solar System, relative to a hydrogen- parable to most bodies in the Solar System, including
rich gas of solar composition, is unresolved. It is pos- Earth. However, ∼1/4 of the studied WDs do not yield
sible that different processes are involved at different lower limits in oxygen fugacity, implying more reduced
stages of planet formation. During rock formation, the parent bodies may be recorded by a minority of sampled
presence of water could increase fO2 in dust that even- objects.
tually forms planetesimals. After rock formation but Intrinsic oxygen fugacity is an important factor that
prior to planetary assembly, aqueous alteration in the influences fundamental characteristics of a planet, in-
parent body could further increase fO2 as iron metal is cluding its volatile budget, its capacity to generate a
oxidized to form iron oxides and iron-bearing clay min- magnetic field, and its prospects for plate tectonics. The
erals. Finally, the addition of Si and/or O to the metal oxygen fugacities of rocky bodies in the Solar System ex-
core of a planetary embryo or planet would result in the hibit a range of ∆IW values of ∼ 7 dex. Earth is indeed
partitioning of Fe into the rock as FeO, and would also oxidized relative to solar, but it is also the most reduced
18 Doyle et al.

of these rocky bodies excluding Mercury and enstatite accretion will come greater confidence in the geochemi-
chondrites. We conclude that to build an Earth-like cal parameters derived from polluted WDs. Such eval-
planet, a reservoir of relatively reduced rocky material uations will be the focus of future work. Results will
is required. The ability to detect an extrasolar Mercury shape future hypotheses for rock formation in the early
would confirm that such reservoirs are available. Solar System and the favored environments for planet
The difficulty in constraining the oxidation states of formation in general.
these more reduced bodies is discussed and a model for
the time-dependent evolution of the apparent oxygen ACKNOWLEDGEMENTS
fugacity for a hypothetical reduced body consumed by This work was supported by NASA 2XRP grant no.
a WD is investigated. Differences in diffusive fluxes of 80NSSC20K0270 to EDY. HES gratefully acknowledges
various elements through the WD envelope yield spuri- support from the National Aeronautics and Space Ad-
ous inferred bulk elemental compositions and oxidation ministration under grant No. 17 NAI18 2 − 0029 is-
states of the accreting parent bodies under certain condi- sued through the NExSS Program. We would like to
tions. The worst case for biasing against detection of re- thank the anonymous referee for their comments which
duced bodies occurs for high effective temperatures. For improved the manuscript.
moderate and low effective temperatures, evidence for
relatively reduced parent bodies is preserved under most
circumstances for at least several characteristic lifetimes
of the debris disk.
In general, detailed understanding of both disk life-
times around WDs and accretion times are important
for assessing exoplanetary oxidation states. In order for
the current method to unambiguously detect a low oxy-
gen fugacity parent body in a WD, the white dwarfs
would need to be accreting mainly the silicate portion of
the body without large mass fractions of metal. In view
of the uncertainties, and the importance of characteriz-
ing the range of extrasolar rock oxidation states, more
attention should be given to the five WDs where lower
limits in xrock
FeO are consistent with zero in this work. Our
calculations suggest that these are candidates for accre-
tion of relatively low oxidation state parent bodies.
While reduced bodies in the Solar System are less com-
mon, the difficulty in constraining them in extrasolar
planetary systems will potentially bias our observations
against observing them among extrasolar rocks. There-
fore, more observations, especially of white dwarfs with
debris disks, is the best way to obtain a better statistical
characterization of extrasolar rock ∆IW values. Strate-
gies for reducing uncertainties in the observations and
the extraction of element ratios from the data are also
important. We estimate that roughly half the errors
in element ratios derive from the raw data themselves
and the other half from dispersion in model results for a
given star. The strategies may therefore involve longer
integration times, in some cases, and the removal of cor-
related uncertainties from modeled values (Klein et al.
in prep.).
It is clear that a better understanding of disk lifetimes
is needed in order to draw definitive conclusions about
the veracity of element ratios of accreted parent bod-
ies. With this better understanding of the timescales of
Where are the Extrasolar Mercuries? 19

APPENDIX

Here we show the derivation for the exponential decay of a debris disk accreting onto a white dwarf simultaneous
with settling of material through the white dwarf atmosphere, as in Jura et al. (2009). We start with the expression
for mass balance of element z in the convective layer of the WD (e.g. Dupuis et al. 1993; Koester 2009):

dMCV,z (t) MCV,z (t)


= Ṁz (t) − , (1)
dt τz
where t is the elapsed time for accretion, MCV,z (t) is the mass abundance of element z in the convection zone, Ṁz (t)
is the rate of accretion of element z onto the star, and τz is the diffusion timescale for element z . The first term on the
right side is the rate of accretion of element z onto the star and the second term is the rate at which element z leaves
the convection zone due to gravitational settling. We replace Ṁz (t) with a time variable accretion rate that depends
on the mass of the disk at time t and a fixed characteristic disk lifetime τdisk :

MPB,z e−t/τdisk
Ṁz (t) = . (2)
τdisk
Substituting Equation 2 into Equation 1 gives

dMCV,z (t) MPB,z e−t/τdisk MCV,z (t)


= − , (3)
dt τdisk τz
which is a first-order differential equation with the integrating factor et/τz . Multiplying both sides by the integrating
factor yields

dMCV,z (t) MCV,z (t) MPB,z e−t/τdisk et/τz


et/τz + et/τz = . (4)
dt τz τdisk
By the product rule the left-hand side of Equation 4 is seen to be (et/τz MCV,z )0 , allowing the simplification

MPB,z t(1/τz −1/τdisk )


(et/τz MCV,z )0 = e . (5)
τdisk
Integration of both sides yields

e(t/τz −t/τdisk )
 
MPB,z
et/τz MCV,z = + c. (6)
τdisk (1/τz − 1/τdisk )
Solving for MCV,z
 (t/τz −t/τdisk ) 
MPB,z 1 e c
MCV,z = t/τ
+ t/τ (7)
τdisk (1/τz − 1/τdisk ) e z e z
and simplifying we obtain

MPB,z 1 c
MCV,z = e−t/τdisk + t/τ . (8)
τdisk (1/τz − 1/τdisk ) e z
Under the initial condition that MCV,z = 0 at t = 0,

MPB,z 1 c
0= e0 + 0 , (9)
τdisk (1/τz − 1/τdisk ) e
and the constant of integration is seen to be

MPB,z 1
c = (−1) . (10)
τdisk (1/τz − 1/τdisk )
20 Doyle et al.

Our solution is therefore

MPB,z 1 MPB,z 1
MCV,z = e−t/τdisk − e−t/τz
τdisk (1/τz − 1/τdisk ) τdisk (1/τz − 1/τdisk )
MPB,z 1 h i
= e−t/τdisk − e−t/τz (11)
τdisk (1/τz − 1/τdisk )
MPB,z 1 τz τdisk h −t/τdisk i
= e − e−t/τz
τdisk (1/τz − 1/τdisk ) τz τdisk

which simplifies to the form shown in Jura et al. (2009):

MPB,z τz h −t/τdisk i
MCV,z = e − e−t/τz . (12)
τdisk − τz

REFERENCES
Badro, J., Brodholt, J. P., Piet, H., Siebert, J., & Ryerson, Dupuis, J., Fontaine, G., Pelletier, C., & Wesemael, F.
F. J. 2015, Proceedings of the National Academy of 1993, The Astrophysical Journal Supplement Series, 84,
Sciences, 112, 12310, doi: 10.1073/pnas.1505672112 73, doi: 10.1086/191746
Barber, S. D., Patterson, A. J., Kilic, M., et al. 2012, Elkins-Tanton, L. T., & Seager, S. 2008, Astrophysical
Astrophysical Journal Journal, 688, 628. hGotoISIi://WOS:000260769700050
Bauer, E. B., & Bildsten, L. 2018, The Astrophysical Farihi, J. 2009, Monthly Notices of the Royal Astronomical
Journal, 859, L19, doi: 10.3847/2041-8213/aac492 Society
—. 2019, The Astrophysical Journal, 872, 96, —. 2016, New Astronomy Reviews, 71, 934,
doi: 10.3847/1538-4357/ab0028 doi: 10.1016/j.newar.2016.03.001
Becklin, E. E., Farihi, J., Jura, M., et al. 2005, Farihi, J., Brinkworth, C. S., Gansicke, B. T., et al. 2011,
Astrophys.J., 632, L119 Astrophysical Journal Letters, 728,
Buffett, B. 2000, Science, 288, 2007 doi: ArtnL810.1088/2041-8205/728/1/L8
Cartier, C., & Wood, B. J. 2019, Elements, 15, 39
Farihi, J., Gansicke, B. T., & Koester, D. 2013, Science,
Chau, A., Reinhardt, C., Helled, R., & Stadel, J. 2018, The
342, 218.
Astrophysical Journal, 865, 35,
https://www.ncbi.nlm.nih.gov/pubmed/24115434
doi: 10.3847/1538-4357/aad8b0
Farihi, J., Gänsicke, B. T., Wyatt, M. C., et al. 2012,
Cunningham, T., Tremblay, P.-E., Freytag, B., Ludwig,
Monthly Notices of the Royal Astronomical Society, 424,
H.-G., & Koester, D. 2019, Monthly Notices of the Royal
464, doi: 10.1111/j.1365-2966.2012.21215.x
Astronomical Society, 488, 25032522,
Farihi, J., Koester, D., Zuckerman, B., et al. 2016, Monthly
doi: 10.1093/mnras/stz1759
Notices of the Royal Astronomical Society, 463, 3186,
Deal, M., Deheuvels, S., Vauclair, G., Vauclair, S., &
doi: 10.1093/mnras/stw2182
Wachlin, F. C. 2013, Astronomy & Astrophysics, 557,
L12, doi: 10.1051/0004-6361/201322206 Fortin-Archambault, M., Dufour, P., & Xu, S. 2020, The
Dennihy, E., Farihi, J., Fusillo, N. P. G., & Debes, J. H. Astrophysical Journal, 888, 47,
2020, Astrophysical Journal doi: 10.3847/1538-4357/ab585a
Dorn, C., Khan, A., Heng, K., et al. 2015, A&A, 577, A83, Frost, D. J., Langenhorst, F., & van Aken, P. A. 2001,
doi: 10.1051/0004-6361/201424915 Physics and Chemistry of Minerals, 28, 455,
Doyle, A. E., Young, E. D., Klein, B., Zuckerman, B., & doi: 10.1007/s002690100181
Schlichting, H. E. 2019, Science, 366, 356, Frost, D. J., & McCammon, C. A. 2008, Annual Review of
doi: 10.1126/science.aax3901 Earth and Planetary Sciences, 36, 389,
Dufour, P., Kilic, M., Fontaine, G., et al. 2012, The doi: 10.1146/annurev.earth.36.031207.124322
Astrophysical Journal, 749, 15 Gansicke, B. T., Koester, D., Farihi, J., et al. 2012,
Dupuis, J., Fontaine, G., Pelletier, C., & Wesemael, F. 1992, Monthly Notices of the Royal Astronomical Society, 424,
The Astrophysical Journal, 82, 505, doi: 10.1086/191728 333, doi: 10.1111/j.1365-2966.2012.21201.x
Where are the Extrasolar Mercuries? 21

Gänsicke, B. T., Marsh, T. R., Southworth, J., & Krot, A., Fegley, B., & Lodders, K. 2000, in Protostars and
Rebassa-Mansergas, A. 2006, Science, 314, 1908, Planets IV, ed. V. Mannings, A. P. Boss, & S. S. Russell
doi: 10.1126/science.1135033 (University of Arizona Press), 1019–1054
Girven, J., Brinkworth, C. S., Farihi, J., et al. 2012, The Kupka, F., Zaussinger, F., & Montgomery, M. H. 2018,
Astrophysical Journal, 749, 11 Monthly Notices of the Royal Astronomical Society, 474,
Grossman, L., Fedkin, A. V., & Simon, S. B. 2012, 4660, doi: 10.1093/mnras/stx3119
Meteoritics & Planetary Science, 47, 2160, Lodders, K. 2003, The Astrophysical Journal, 591, 1220,
doi: 10.1111/j.1945-5100.2012.01353.x doi: 10.1086/375492
Gnsicke, B. T., Schreiber, M. R., Toloza, O., et al. 2019, Manser, C. J., Gnsicke, B. T., GentileFusillo, N. P., et al.
Nature, 576, 6164, doi: 10.1038/s41586-019-1789-8 2020, Monthly Notices of the Royal Astronomical
Hollands, M. A., Gänsicke, B. T., & Koester, D. 2018, Society, 493, 21272139, doi: 10.1093/mnras/staa359
Monthly Notices of the Royal Astronomical Society, 477, Manser, C. J., Gänsicke, B. T., Eggl, S., et al. 2019,
93, doi: 10.1093/mnras/sty592 Science, 364, 66, doi: 10.1126/science.aat5330
Jura, M. 2003, Astrophysical Journal, 584, L91 McDonough, W. F. 2003, in Treatise on Geochemistry, ed.
—. 2008, The Astronomical Journal, 135, 1785, K. K. Turekian & H. D. Holland (Elsevier), 547–568,
doi: 10.1088/0004-6256/135/5/1785 doi: 10.1016/b0-08-043751-6/02015-6
Jura, M., Farihi, J., & Zuckerman, B. 2007, The Melis, C., & Dufour, P. 2017, The Astrophysical Journal,
834, doi: 10.3847/1538-4357/834/1/1
Astrophysical Journal, 663, 1285, doi: 10.1086/518767
Melis, C., Jura, M., Albert, L., Klein, B., & Zuckerman, B.
Jura, M., Muno, M. P., Farihi, J., & Zuckerman, B. 2009,
2010, The Astrophysical Journal, 722, 1078,
Astrophys.J., 699, 1473
doi: 10.1088/0004-637x/722/2/1078
Jura, M., & Xu, S. 2010, The Astronomical Journal, 140,
Monteux, J., Golabek, G. J., Rubie, D. C., Tobie, G., &
1129
Young, E. D. 2018, Space Science Reviews, 214,
—. 2012, The Astronomical Journal, 143:6,
doi: 10.1007/s11214-018-0473-x
doi: 10.1088/0004-6256/143/1/6
Newsom, H. E. 1995, AGU Reference Shelf, Vol. 1,
Jura, M., Xu, S., Klein, B., Koester, D., & Zuckerman, B.
Composition of the solar system, planets, meteorites, and
2012, Astrophysical Journal, 750.
major terrestrial reservoirs, 159–189,
hGotoISIi://WOS:000303063500069
doi: 10.1029/RF001p0159
Jura, M., & Young, E. D. 2014, Annual Review of Earth
Nittler, L. R., Chabot, N. L., Grove, T. L., & Peplowski,
and Planetary Sciences, Vol 42, 42, 45,
P. N. 2019, The Chemical Composition of Mercury, ed.
doi: 10.1146/annurev-earth-060313-054740
N. . A. Solomon (Cambridge Univerity Press)
Kilic, M., Farihi, J., Nitta, A., & Leggett, S. K. 2008, The
Pearson, D. G., Brenker, F. E., Nestola, F., et al. 2014,
Astronomical Journal, 136, 111,
Nature, 507, 221, doi: 10.1038/nature13080
doi: 10.1088/0004-6256/136/1/111 Raddi, R., Gansicke, B. T., Koester, D., et al. 2015,
Klein, B., Jura, M., Koester, D., & Zuckerman, B. 2011, Monthly Notices of the Royal Astronomical Society, 450,
The Astrophysical Journal, 741, 2083, doi: 10.1093/mnras/stv701
doi: 10.1088/0004-637x/741/1/64 Rafikov, R. R. 2011a, The Astrophysical Journal, 732, L3,
Klein, B., Jura, M., Koester, D., Zuckerman, B., & Melis, doi: 10.1088/2041-8205/732/1/l3
C. 2010, Astrophysical Journal, 709, 950, —. 2011b, Monthly Notices of the Royal Astronomical
doi: 10.1088/0004-637x/709/2/950 Society: Letters, 416, L55L59,
Koester, D. 2009, Astronomy & Astrophysics, 498, 517, doi: 10.1111/j.1745-3933.2011.01096.x
doi: 10.1051/0004-6361/200811468 Rubie, D. C., Frost, D. J., Mann, U., et al. 2011, Earth and
—. 2013. http://www1.astrophysik.uni-kiel.de/koester/ Planetary Science Letters, 301, 31
astrophysics/astrophysics.html Rubie, D. C., Jacobson, S. A., Morbidelli, A., et al. 2015,
Koester, D., Gansicke, B. T., & Farihi, J. 2014, Astronomy Icarus, 248, 89
& Astrophysics, 566, Schaefer, L., & Fegley, B. 2017, The Astrophysical Journal,
doi: ArtnA3410.1051/0004-6361/201423691 843, 120, doi: 10.3847/1538-4357/aa784f
Kolzenburg, S., Di Genova, D., Giordano, D., Hess, K. U., Schrder, K.-P., & Connon Smith, R. 2008, Monthly Notices
& Dingwell, D. B. 2018, Earth and Planetary Science of the Royal Astronomical Society, 386, 155,
Letters, 487, 21, doi: 10.1016/j.epsl.2018.01.023 doi: 10.1111/j.1365-2966.2008.13022.x
22 Doyle et al.

Steenstra, E., Knibbe, J., Rai, N., & van Westrenen, W. Wyatt, M. C., Farihi, J., Pringle, J. E., & Bonsor, A. 2014,
2016, Geochemica et Cosmochimica Acta, 177, 48 Monthly Notices of the Royal Astronomical Society, 439,
Swan, A., Farihi, J., Koester, D., et al. 2019, Monthly 33713391, doi: 10.1093/mnras/stu183
Notices of the Royal Astronomical Society Xu, S., Dufour, P., Klein, B., et al. 2019, The Astronomical
Taylor, G. J. 2013, Chemie der Erde - Geochemistry, 73, Journal, 158
401, doi: 10.1016/j.chemer.2013.09.006 Xu, S., Jura, M., Koester, D., Klein, B., & Zuckerman, B.
Tremblay, P. E., Ludwig, H. G., Freytag, B., Koester, D., & 2014, Astrophysical Journal
Fontaine, G. 2017, Mem. Soc. Astron. Italiana, 88, 104 Xu, S., Zuckerman, B., Dufour, P., et al. 2017, The
Vanderbosch, Z., Hermes, J. J., Dennihy, E., et al. 2019, Astrophysical Journal, 836,
Astrophysical Journal Letters doi: 10.3847/2041-8213/836/1/l7
Vanderburg, A., Johnson, J. A., Rappaport, S., et al. 2015, Zuckerman, B., & Becklin, E. E. 1987, Nature, 330, 138,
Nature, 526, 546 doi: 10.1038/330138a0
Veras, D. 2016, Post-main-sequence planetary system Zuckerman, B., Koester, D., Reid, I. N., & Hunsch, M.
evolution. https://arxiv.org/abs/1601.05419 2003, Astrophysical Journal, 596, 477,
Wade, J., & Wood, B. 2005, Earth and Planetary Science doi: Doi10.1086/377492
Letters, 236, 78 Zuckerman, B., Melis, C., Klein, B., Koester, D., & Jura,
Wilson, D. J., Gaensicke, B. T., Koester, D., et al. 2015, M. 2010, Astrophysical Journal, 722, 725,
Monthly Notices of the Royal Astronomical Society, 451, doi: 10.1088/0004-637x/722/1/725
3237, doi: 10.1093/mnras/stv1201 Zuckerman, B., & Young, E. D. 2017, Handbook of
Wood, B. J., Walter, M. J., & Wade, J. 2006, Nature, 441, Exoplanets, doi: 10.1007/978-3-319-30648-3 14-1
825

You might also like